[go: up one dir, main page]

 
 
remotesensing-logo

Journal Browser

Journal Browser

Advances in Quantitative Remote Sensing in China – In Memory of Prof. Xiaowen Li

A special issue of Remote Sensing (ISSN 2072-4292).

Deadline for manuscript submissions: closed (7 January 2018) | Viewed by 246566

Printed Edition Available!
A printed edition of this Special Issue is available here.

Special Issue Editors


E-Mail Website
Guest Editor
Department of Geographical Sciences, University of Maryland, College Park, MD 20742, USA
Interests: quantitative land remote sensing; Earth’s energy budget; global environmental change
Special Issues, Collections and Topics in MDPI journals

E-Mail
Guest Editor
Institute of Remote Sensing Science and Engineering, Faculty of Geographical Science, Beijing Normal University, Beijing 100875, China
Interests: multiangular remote sensing; vegetation remote sensing; radiation budget

E-Mail
Guest Editor
Institute of Remote Sensing and Digital Earth, CAS, China
Interests: microwave remote sensing of water cycle related components

Special Issue Information

Dear Colleagues,

China has recently developed a very comprehensive and ambitious Earth observation program. Hundreds of satellites have been launched or will be lunched soon. A series of grand research projects have been funded to process and analyze the huge amount of satellite data. As a result, significant progress in quantitative remote sensing has been made from radiative transfer modeling, advanced inversion methods, high-level products generation, to various applications.

In memory of Prof. Xiaowen Li, who was one of the pioneers in promoting quantitative remote sensing study in China, we organized the third national forum on quantitative remote sensing in Beijing Normal University, July 14 and 15, 2017. More than 200 people attended this forum.

To document the progress and facilitate more international collaborations, we propose to edit this Special Issue on Remote Sensing.

It will cover the full aspects of quantitative remote sensing, with particular focus on:

  • New satellite missions
  • Radiative transfer modeling
  • Inversion methodology
  • Satellite products generation
  • Field measurements and validation
  • Satellite product applications

In addition to regular submissions, some review papers will be invited. We will also write an Editorial to summarize the achievements of Prof. Xiaowen Li as a memorial.

Thank you for your consideration.

Dr. Shunlin Liang
Dr. Guangjian Yan
Dr. Jiancheng Shi
Guest Editors

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. Remote Sensing is an international peer-reviewed open access semimonthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. The Article Processing Charge (APC) for publication in this open access journal is 2700 CHF (Swiss Francs). Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Benefits of Publishing in a Special Issue

  • Ease of navigation: Grouping papers by topic helps scholars navigate broad scope journals more efficiently.
  • Greater discoverability: Special Issues support the reach and impact of scientific research. Articles in Special Issues are more discoverable and cited more frequently.
  • Expansion of research network: Special Issues facilitate connections among authors, fostering scientific collaborations.
  • External promotion: Articles in Special Issues are often promoted through the journal's social media, increasing their visibility.
  • e-Book format: Special Issues with more than 10 articles can be published as dedicated e-books, ensuring wide and rapid dissemination.

Further information on MDPI's Special Issue polices can be found here.

Published Papers (41 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Editorial

Jump to: Research, Review

7 pages, 558 KiB  
Editorial
Recent Progress in Quantitative Land Remote Sensing in China
by Shunlin Liang, Jiancheng Shi and Guangjian Yan
Remote Sens. 2018, 10(9), 1490; https://doi.org/10.3390/rs10091490 - 18 Sep 2018
Cited by 4 | Viewed by 4119
Abstract
During the past forty years, since the first book with a title mentioning quantitative and remote sensing was published [1], quantitative land remote sensing has advanced dramatically, and numerous books have been published since then [26] although [...] Read more.
During the past forty years, since the first book with a title mentioning quantitative and remote sensing was published [1], quantitative land remote sensing has advanced dramatically, and numerous books have been published since then [26] although some of them did not use quantitative land remote sensing in their titles. [...]
Full article
Show Figures

Figure 1

Figure 1
<p>The scope of quantitative land remote sensing.</p>
Full article ">Figure 2
<p>Major Chinese satellites relevant to land remote sensing.</p>
Full article ">

Research

Jump to: Editorial, Review

11 pages, 877 KiB  
Article
From Geometric-Optical Remote Sensing Modeling to Quantitative Remote Sensing Science—In Memory of Academician Xiaowen Li
by Qinhuo Liu, Guangjian Yan, Ziti Jiao, Qing Xiao, Jianguang Wen, Shunlin Liang, Jindi Wang, Crystal Schaaf and Alan Strahler
Remote Sens. 2018, 10(11), 1764; https://doi.org/10.3390/rs10111764 - 8 Nov 2018
Cited by 8 | Viewed by 5952
Abstract
The academician Xiaowen Li devoted much of his life to pursuing fundamental research in remote sensing. A pioneer in the geometric-optical modeling of vegetation canopies, his work is held in high regard by the international remote sensing community. He codeveloped the Li–Strahler geometric-optic [...] Read more.
The academician Xiaowen Li devoted much of his life to pursuing fundamental research in remote sensing. A pioneer in the geometric-optical modeling of vegetation canopies, his work is held in high regard by the international remote sensing community. He codeveloped the Li–Strahler geometric-optic model, and this paper was selected by a member of the International Society for Optical Engineering (SPIE) milestone series. As a chief scientist, Xiaowen Li led a scientific team that made outstanding advances in bidirectional reflectance distribution modeling, directional thermal emission modeling, comprehensive experiments, and the understanding of spatial and temporal scale effects in remote sensing information, and of quantitative inversions utilizing remote sensing data. In addition to his broad research activities, he was noted for his humility and his dedication in making science more accessible for the general public. Here, the life and academic contributions of Xiaowen Li to the field of quantitative remote sensing science are briefly reviewed. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Academician Xiaowen Li.</p>
Full article ">
25 pages, 2731 KiB  
Article
Effects of Forest Canopy Vertical Stratification on the Estimation of Gross Primary Production by Remote Sensing
by Shangrong Lin, Jing Li, Qinhuo Liu, Alfredo Huete and Longhui Li
Remote Sens. 2018, 10(9), 1329; https://doi.org/10.3390/rs10091329 - 21 Aug 2018
Cited by 15 | Viewed by 5465
Abstract
Gross primary production (GPP) in forests is the most important carbon flux in terrestrial ecosystems. Forest ecosystems with high leaf area index (LAI) values have diverse species or complex forest structures with vertical stratifications that influence the carbon–water–energy cycles. In this study, we [...] Read more.
Gross primary production (GPP) in forests is the most important carbon flux in terrestrial ecosystems. Forest ecosystems with high leaf area index (LAI) values have diverse species or complex forest structures with vertical stratifications that influence the carbon–water–energy cycles. In this study, we used three light use efficiency (LUE) GPP models and site-level experiment data to analyze the effects of the vertical stratification of dense forest vegetation on the estimates of remotely sensed GPP during the growing season of two forest sites in East Asia: Dinghushan (DHS) and Tomakomai (TMK). The results showed that different controlling environmental factors of the vertical layers, such as temperature and vapor pressure deficit (VPD), produce different responses for the same LUE value in the different sub-ecosystems (defined as the tree, shrub, and grass layers), which influences the GPP estimation. Air temperature and VPD play important roles in the effects of vertical stratification on the GPP estimates in dense forests, which led to differences in GPP uncertainties from −50% to 30% because of the distinct temperature responses in TMK. The unequal vertical LAI distributions in the different sub-ecosystems led to GPP variations of 1–2 gC/m2/day with uncertainties of approximately −30% to 20% because sub-ecosystems have unique absorbed fractions of photosynthetically active radiation (APAR) and LUE. A comparison with the flux tower-based GPP data indicated that the GPP estimations from the LUE and APAR values from separate vertical layers exhibited better model performance than those calculated using the single-layer method, with 10% less bias in DHS and more than 70% less bias in TMK. The precision of the estimated GPP in regions with thick understory vegetation could be effectively improved by considering the vertical variations in environmental parameters and the LAI values of different sub-ecosystems as separate factors when calculating the GPP of different components. Our results provide useful insight that can be used to improve the accuracy of remote sensing GPP estimations by considering vertical stratification parameters along with the LAI of sub-ecosystems in dense forests. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Locations of the two study sites in East Asia.</p>
Full article ">Figure 2
<p>Seasonal patterns in the vertical variance in temperature and vapor pressure deficit (VPD) at the DHS and TMK sites. In the subplots (<b>A,C</b>), the blue bars (VPDdif in Pa) are the biases between the tree layer VPD and the grass layer VPD. The red line (Tdif in °C) is the bias between the air temperatures of the top and grass layers. The green dotted line is the 0 °C line, indicating no difference between the temperatures of the tree and grass layers. Subplots (<b>B,D</b>) show the relationships between Tdif and VPDdif at the two sites. The dot colors are clustered by month.</p>
Full article ">Figure 3
<p>Bar chart for the daily average incident PAR in sub-ecosystems at the selected sites; each bar represents the daily mean incident PAR in each layer (<b>A</b>) TMK; (<b>B</b>) DHS.</p>
Full article ">Figure 4
<p>Modeled LUE values in the different sub-ecosystems.</p>
Full article ">Figure 5
<p>Differences in gross primary production (GPP) under different mean canopy temperatures. The GPP difference is calculated as (GPPm-GPPs)/GPPs. DHS_VPM, DHS_W, and DHS_MOD are the values calculated at the DHS site and by VPM, Wu’s method, and the MOD17 model, respectively.</p>
Full article ">Figure 6
<p>Effects of both considering and not considering the vertical differences in VPD on GPP. The black line is the fit line for the GPP difference scatterplot at different sites. Subfigures (<b>A</b>,<b>B</b>) are the GPP difference comparisons based on different canopy mean VPDs at TMK and DHS, respectively.</p>
Full article ">Figure 7
<p>Ternary contour for the differences between considering and not considering the vertical structure during the GPP estimation with LAI separated into tree, shrub, and grass layers. This graph illustrates that when the total LAI of the canopy is four, there is little temperature and humidity stress on the LUE. We assumed that the LUE of the grass layer was 1.28 gC/m<sup>2</sup>/day/MJ, that of the shrub layer was 0.85 gC/m<sup>2</sup>/day/MJ, and that of the tree layer was 1.02 gC/m<sup>2</sup>/day/MJ.</p>
Full article ">Figure 8
<p>GPP differences considering multilayer and single-layer estimations of incident PAR. Subfigure (<b>A</b>,<b>C</b>) show the GPP simulation results with flux tower measurements of GPP. The green points represent multilayer results (GPPmMOD), and the red points represent single-layer results (GPPsMOD). The black points represent the flux tower-based GPP results (GPPactual). Subfigure (<b>B</b>,<b>D</b>) show GPP differences under various incident PAR.</p>
Full article ">Figure 9
<p>Differences in GPP estimations in different seasons. Subfigure (<b>A</b>) showed the seasonal GPP variance in DHS, while subfigure (<b>B</b>) showed the condition in TMK. Here, we used three different types of LUEmax parametric approaches to model GPP using Wu’s method. GPPmW is the multilayer result using two or three different LUEmax inputs for GPP estimation. GPPsW is the single-layer result using an EBF-based LUEmax. GPPaW is the average LUEmax input in GPPmW, using the average LUEmax for the different components.</p>
Full article ">Figure 9 Cont.
<p>Differences in GPP estimations in different seasons. Subfigure (<b>A</b>) showed the seasonal GPP variance in DHS, while subfigure (<b>B</b>) showed the condition in TMK. Here, we used three different types of LUEmax parametric approaches to model GPP using Wu’s method. GPPmW is the multilayer result using two or three different LUEmax inputs for GPP estimation. GPPsW is the single-layer result using an EBF-based LUEmax. GPPaW is the average LUEmax input in GPPmW, using the average LUEmax for the different components.</p>
Full article ">Figure 10
<p>Differences in GPP estimation using different models with carbon flux tower measurements. Subfigure (<b>a</b>) showed the condition in DHS, and subfigure (<b>b</b>) showed it in TMK. VPMm results from using the multilayer VPM GPP minus the flux tower-based GPP. VPMs results from using the single-layer VPM GPP minus the flux tower-based GPP. The Y-axis on the right is the daily mean shortwave incident radiation.</p>
Full article ">Figure 11
<p>Vertical LAI ratios in different ecosystems (EBF, evergreen broadleaf forest [<a href="#B71-remotesensing-10-01329" class="html-bibr">71</a>,<a href="#B90-remotesensing-10-01329" class="html-bibr">90</a>]; DNF, deciduous needleleaf forest [<a href="#B72-remotesensing-10-01329" class="html-bibr">72</a>]; ENF, evergreen needleleaf forest [<a href="#B91-remotesensing-10-01329" class="html-bibr">91</a>,<a href="#B92-remotesensing-10-01329" class="html-bibr">92</a>]; DBF, deciduous broadleaf forest [<a href="#B26-remotesensing-10-01329" class="html-bibr">26</a>]; SAV, savanna [<a href="#B93-remotesensing-10-01329" class="html-bibr">93</a>]).</p>
Full article ">
22 pages, 19325 KiB  
Article
The Retrieval of 30-m Resolution LAI from Landsat Data by Combining MODIS Products
by Jianmin Zhou, Shan Zhang, Hua Yang, Zhiqiang Xiao and Feng Gao
Remote Sens. 2018, 10(8), 1187; https://doi.org/10.3390/rs10081187 - 27 Jul 2018
Cited by 9 | Viewed by 5380
Abstract
Leaf area index (LAI) is a critical vegetation structural parameter in biogeochemical and biophysical ecosystems. High-resolution LAI products play an essential role in regional studies. Empirical methods, which normally use field measurements as their training samples and have been identified as the most [...] Read more.
Leaf area index (LAI) is a critical vegetation structural parameter in biogeochemical and biophysical ecosystems. High-resolution LAI products play an essential role in regional studies. Empirical methods, which normally use field measurements as their training samples and have been identified as the most commonly used approaches to retrieve structural parameters of vegetation from high-resolution remote-sensing data, are limited by the quality of training samples. Few efforts have been made to generate training samples from existing global LAI products. In this study, two methods (a homogeneous and pure pixel filter method (method A) and a pixel unmixing method (method B)) were developed to extract training samples from moderate-resolution imaging spectroradiometer (MODIS) surface reflectance and LAI products, and a support vector regression (SVR) algorithm trained by the samples was used to retrieve the high-resolution LAI from Landsat data at Baoding, situated in the Hebei Province in China, and Des Moines, situated in Iowa, United States. For the homogeneous and pure pixel filter method, two different sets of training samples were designed. One was composed of upscaled Landsat reflectance at the 500-m resolution and MODIS LAI products (dataset A1); the other was composed of MODIS reflectance and LAI products (dataset A2). With them, two inversion models were developed using SVR. For the pixel unmixing method, the training samples (dataset B) were extracted from unmixed MODIS surface reflectance and LAI products at 30-m resolution, and the third inversion model was obtained with them. LAI inversion results showed that good agreement with field measurements was achieved using these three inversion models. The R2 (coefficient of determination) value and the root mean square error (RMSE) value were computed to assess the results. For all tests, the R2 values are higher than 0.74 and RMSE values are less than 0.73. These tests showed that three models for the two methods combined with MODIS products can retrieve 30-m resolution LAI from Landsat data. The results of the pixel unmixing method was slightly better than that of the homogeneous and pure pixel filter method. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Data processing flowchart of the support vector regression (SVR) inversion method with two training sample-selecting methods: (<b>A</b>) the homogeneous and pure pixel filter method and (<b>B</b>) the pixel unmixing method.</p>
Full article ">Figure 2
<p>The two study areas: (<b>a</b>) Baoding, Hebei Province, China; (<b>b</b>) Des Moines, Iowa, United States. The grid in (<b>b</b>) shows the MODIS pixel cells at the 500-m resolution.</p>
Full article ">Figure 2 Cont.
<p>The two study areas: (<b>a</b>) Baoding, Hebei Province, China; (<b>b</b>) Des Moines, Iowa, United States. The grid in (<b>b</b>) shows the MODIS pixel cells at the 500-m resolution.</p>
Full article ">Figure 3
<p>The MODIS leaf area index (LAI) pixel filtering process. From left to right: (<b>a</b>) original MODIS LAI; (<b>b</b>) MODIS LAI pixels for cropland with the highest retrieval quality (main algorithm and not saturated); (<b>c</b>) CV map of NIR band4 from Landsat5; and (<b>d</b>) final selected homogeneous and pure MODIS LAI pixels.</p>
Full article ">Figure 4
<p>Classification products of the Soil Moisture Experiment 2002 (SMEX02) near Des Moines, Iowa, United States: (<b>a</b>) MCD12Q1_IGBP; (<b>b</b>) GlobeLand30 in 2000; (<b>c</b>) GlobeLand30 in 2010; (<b>d</b>) 30-m classification product. Notes: (<b>a</b>) 12: agricultural land; 13: urban and construction area; 14: junction of agricultural land and natural vegetation. (<b>b</b>–<b>d</b>) 10: arable land; 20: forest; 30: grassland; 50: wetland; 60: water body; 80: artificial surfaces; 90: bare land.</p>
Full article ">Figure 4 Cont.
<p>Classification products of the Soil Moisture Experiment 2002 (SMEX02) near Des Moines, Iowa, United States: (<b>a</b>) MCD12Q1_IGBP; (<b>b</b>) GlobeLand30 in 2000; (<b>c</b>) GlobeLand30 in 2010; (<b>d</b>) 30-m classification product. Notes: (<b>a</b>) 12: agricultural land; 13: urban and construction area; 14: junction of agricultural land and natural vegetation. (<b>b</b>–<b>d</b>) 10: arable land; 20: forest; 30: grassland; 50: wetland; 60: water body; 80: artificial surfaces; 90: bare land.</p>
Full article ">Figure 5
<p>The comparison between the retrieved LAI by three approaches and field measurements: (<b>a</b>) the results of dataset A1 (upscaled Landsat reflectance at the 500-m resolution and MODIS LAI products); (<b>b</b>) the results of dataset A2 (MODIS reflectance and LAI products); (<b>c</b>) the results of dataset B (unmixed MODIS surface reflectance and LAI products at 30-m resolution).</p>
Full article ">Figure 6
<p>Comparison of inversion LAI in time series with field measurements. (<b>a</b>–<b>c</b>) The results of dataset A1; (<b>d</b>–<b>f</b>) the results of dataset A2; (<b>g</b>–<b>i</b>) the results of dataset B.</p>
Full article ">Figure 7
<p>Comparison of inversion LAI with field measurements at the Baoding study area; (<b>a</b>) the results of datasets A1 and A2; (<b>b</b>) the results of dataset B.</p>
Full article ">Figure 8
<p>The LAI inversion in time series at the SMEX02 site (23 June, 1 July, and 8 July): (<b>a</b>) the results of dataset A1; (<b>b</b>) the results of dataset A2; (<b>c</b>) the results of dataset B.</p>
Full article ">Figure 9
<p>LAI inversion in time series at the Baoding area (MODIS DOY: 89–129), based on the homogeneous and pure pixel filter method and the pixel unmixing method. (<b>a</b>) pixel one; (<b>b</b>) pixel two; (<b>c</b>) pixel three; (<b>d</b>) pixel four; (<b>e</b>) pixel five and (<b>f</b>) pixel six.</p>
Full article ">Figure 10
<p>Histogram of LAI from the LAI-SR samples of SMEX02. (<b>a</b>) dataset A1 (upscaled Landsat reflectance at the 500-m resolution and MODIS LAI products); (<b>b</b>) dataset A2 (MODIS reflectance and LAI products); (<b>c</b>) dataset B (unmixed MODIS surface reflectance and LAI products at 30-m resolution).</p>
Full article ">
20 pages, 2707 KiB  
Article
Developing an Integrated Remote Sensing Based Biodiversity Index for Predicting Animal Species Richness
by Jinhui Wu and Shunlin Liang
Remote Sens. 2018, 10(5), 739; https://doi.org/10.3390/rs10050739 - 10 May 2018
Cited by 14 | Viewed by 5732
Abstract
Many remote sensing metrics have been applied in large-scale animal species monitoring and conservation. However, the capabilities of these metrics have not been well compared and assessed. In this study, we investigated the correlation of 21 remote sensing metrics in three categories with [...] Read more.
Many remote sensing metrics have been applied in large-scale animal species monitoring and conservation. However, the capabilities of these metrics have not been well compared and assessed. In this study, we investigated the correlation of 21 remote sensing metrics in three categories with the global species richness of three different animal classes using several statistical methods. As a result, we developed a new index by integrating several highly correlated metrics. Of the 21 remote sensing metrics analyzed, evapotranspiration (ET) had the greatest impact on species richness on a global scale (explained variance: 52%). The metrics with a high explained variance on the global scale were mainly in the energy/productivity category. The metrics in the texture category exhibited higher correlation with species richness at regional scales. We found that radiance and temperature had a larger impact on the distribution of bird richness, compared to their impacts on the distributions of both amphibians and mammals. Three machine learning models (i.e., support vector machine, random forests, and neural networks) were evaluated for metric integration, and the random forest model showed the best performance. Our newly developed index exhibited a 0.7 explained variance for the three animal classes’ species richness on a global scale, with an explained variance that was 20% higher than any of the univariate metrics. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flowchart of the comparison of the remote sensing metrics and development of a multivariate integration index.</p>
Full article ">Figure 2
<p>Geographic distribution of three animal classes. (<b>a</b>) bird richness, (<b>b</b>) amphibian richness, (<b>c</b>) mammal richness, and (<b>d</b>) combined species richness where we assigned mammal richness to the red band, bird richness to the green band, and amphibian richness to the blue band of the image. All other colors show transition zones of mixtures of the different animal classes.</p>
Full article ">Figure 3
<p>The explained variance of remote sensing metrics on species richness. Percentages of variance explained by the regression model was built with different remote sensing metrics. For DHI-min and DHI-cum, univariate linear regression was used to calculate the explained variance. For COR, CV, and DHI-sea, linearization was achieved by taking the log of the data before using univariate linear regression to calculate the explained variance. For the other metrics, polynomial regression was used to calculate the explained variance. Allclasses is the sum of mammal richness, bird richness, and amphibian richness, which was normalized. The measured values were obtained for 10,000 0.1-degree pixels randomly selected from global terrestrial areas. The full names of abbreviations can be found in the <a href="#remotesensing-10-00739-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 4
<p>The explained variance of DHI metrics on mammal richness at latitudinal zones (360° × 10°). For DHI-min and DHI-cum, univariate linear regression was used to calculate the explained variance. For DHI-sea, linearization was achieved by taking the log of the data before using univariate linear regression to calculate the explained variance. The measured values were obtained for 10,000 0.1-degree pixels randomly selected from global terrestrial areas.</p>
Full article ">Figure 5
<p>Geographic distribution of DHI-cum, DHI-min, mammal richness. (<b>a</b>) North America, (<b>b</b>) Argentina and Chile, (<b>c</b>) Amazon.</p>
Full article ">Figure 6
<p>The Akaike information criterion (AIC) values of the three models.</p>
Full article ">Figure 7
<p>The correlation between predicted species richness and in situ species richness. (<b>a</b>) training dataset, (<b>b</b>) prediction dataset. The measured values were obtained for 10,000 0.1-degree pixels randomly selected from terrestrial areas of the globe, and 5000 pixels were randomly selected for training and prediction separately.</p>
Full article ">Figure 8
<p>Distribution of simulated species richness and in situ species richness at latitudinal/longitudinal zones. The data are from a prediction dataset, which has 5000 samples.</p>
Full article ">
18 pages, 5082 KiB  
Article
Reconstruction of Single Tree with Leaves Based on Terrestrial LiDAR Point Cloud Data
by Donghui Xie, Xiangyu Wang, Jianbo Qi, Yiming Chen, Xihan Mu, Wuming Zhang and Guangjian Yan
Remote Sens. 2018, 10(5), 686; https://doi.org/10.3390/rs10050686 - 28 Apr 2018
Cited by 17 | Viewed by 7970
Abstract
Many studies have been focusing on reconstructing the branch skeleton of a three-dimensional (3D) tree structure that is based on photos or point clouds scanned by a terrestrial laser scanner (TLS), but leaves, as the important component of a tree, are often ignored [...] Read more.
Many studies have been focusing on reconstructing the branch skeleton of a three-dimensional (3D) tree structure that is based on photos or point clouds scanned by a terrestrial laser scanner (TLS), but leaves, as the important component of a tree, are often ignored or simplified because of their complexity. Therefore, we develop a voxel-based method to add leaves to a reconstructed 3D branches structure based on TLS point clouds. The location and size of each leaf depend on the spatial distribution and density of leaves points in the voxel. We reconstruct a small 3D scene with four realistic 3D trees and a virtual tree (including trunk, branches, and leaves), and validate the structure of each tree through the directional gap fractions calculated based on simulated point clouds of this reconstructed scene by the ray-tracing algorithm. The results show good coherence with those from measured point clouds data. The relative errors of the directional gap fractions are no more than 4.1%, though the method is limited by the effects of point occlusion. Therefore, this method is shown to give satisfactory consistency both visually and in the quantitative evaluation of the 3D structure. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Relative positions of the four trees (No. 1, 2, 3, and 4) and six scanning sites (orange points), as seen from above. The outlines of the trees were drawn by projecting the scanned points in nadir.</p>
Full article ">Figure 2
<p>Diagram of leaf geometry (<b>a</b>) and its transform (<b>b</b>).</p>
Full article ">Figure 3
<p>Comparison of simulated and real photos: (<b>a</b>) Simulated photograph based on the reconstructed scene; and, (<b>b</b>) Real photograph taken from unmanned aerial vehicle (UAV).</p>
Full article ">Figure 4
<p>(<b>a</b>) A virtual tree generated with OnyxTREE software; (<b>b</b>) The simulated point clouds based on the virtual tree (Blue points represent trunks/branches and red points represent leaves); and, (<b>c</b>) The reconstructed 3D tree.</p>
Full article ">Figure 5
<p>Comparison of the measured (<b>a</b>) and simulated (<b>b</b>) point clouds for one scan. The different colors represent the distances from scanner to objects. The red and blue circles highlight the difference between the measured and simulated point clouds.</p>
Full article ">Figure 6
<p>Comparison of the directional gap fractions from measured and simulated points for each site scan (zenith angle interval: 2.5°).</p>
Full article ">Figure 7
<p>Relationship between the directional gap fractions of measured and simulated points (<b>a</b>) and the distribution of their errors (<b>b</b>).</p>
Full article ">Figure 8
<p>Comparison of the sliced leaf-point densities for each tree (No. 1–4), calculated based on measured and simulated point clouds (interval height: 0.2 m).</p>
Full article ">Figure 9
<p>Relative errors of the gap fractions from the black-and-white images with different view directions simulated based on the original virtual tree and the reconstructed one.</p>
Full article ">Figure 10
<p>Diagram of leaf shape: (<b>a</b>) Square, (<b>b</b>) Rhombus.</p>
Full article ">Figure 11
<p>Comparison of the reconstructed trees using the different leaf shapes, including (<b>a</b>) rhombus, (<b>b</b>) square and (<b>c</b>) typical quadrilateral, with (<b>d</b>) the original virtual tree.</p>
Full article ">Figure 12
<p>Comparison of point densities based on the simulated point clouds of the reconstructed trees with different leaf shapes and the original virtual tree (interval height: 0.2m).</p>
Full article ">
15 pages, 2622 KiB  
Article
Spatial Downscaling of Gross Primary Productivity Using Topographic and Vegetation Heterogeneity Information: A Case Study in the Gongga Mountain Region of China
by Xinyao Xie, Ainong Li, Huaan Jin, Gaofei Yin and Jinhu Bian
Remote Sens. 2018, 10(4), 647; https://doi.org/10.3390/rs10040647 - 22 Apr 2018
Cited by 14 | Viewed by 4064
Abstract
Due to the spatial heterogeneity of land surfaces, downscaling is an important issue in the development of carbon cycle models when evaluating the role of ecosystems in the global carbon cycle. In this study, a downscaling algorithm was developed to model gross primary [...] Read more.
Due to the spatial heterogeneity of land surfaces, downscaling is an important issue in the development of carbon cycle models when evaluating the role of ecosystems in the global carbon cycle. In this study, a downscaling algorithm was developed to model gross primary productivity (GPP) at 500 m in a time series over rugged terrain, which considered the effects of spatial heterogeneity on carbon flux simulations. This work was carried out for a mountainous area with an altitude ranging from 2606 to 4744 m over the Gongga Mountain (Sichuan Province, China). In addition, the Moderate Resolution Imaging Spectroradiometer (MODIS) GPP product at 1 km served as the primary dataset for the downscaling algorithm, and the 500 m MODIS GPP product was used as the reference dataset to evaluate the downscaled GPP results. Moreover, in order to illustrate the advantages and benefits of the proposed downscaling method, the downscaled results in this work, along with ordinary kriging downscaled results, spline downscaled results and inverse distance weighted (IDW) downscaled results, were compared to the MODIS GPP at 500 m. The results showed that (1) the GPP difference between the 500 m MODIS GPP and the proposed downscaled GPP results was primarily in the range of [−1, 1], showing that both vegetation heterogeneity factors (i.e., LAI) and topographic factors (i.e., altitude, slope and aspect) were useful for GPP downscaling; (2) the proposed downscaled results (R2 = 0.89, RMSE = 1.03) had a stronger consistency with the 500 m MODIS GPP than those of the ordinary kriging downscaled results (R2 = 0.43, RMSE = 1.36), the spline downscaled results (R2 = 0.40, RMSE = 1.50) and the IDW downscaled results (R2 = 0.42, RMSE = 1.10) for all Julian days; and (3) the inconsistency between MODIS GPP at 500 m and 1 km increased with the increase in altitude and slope. The proposed downscaling algorithm could provide a reference when considering the effects of spatial heterogeneity on carbon flux simulations and retrieving other fine resolution ecological-physiology parameters (e.g., net primary productivity and evaporation) over topographically complex terrains. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area.</p>
Full article ">Figure 2
<p>Flow chart of the proposed downscaling algorithm.</p>
Full article ">Figure 3
<p>Spatial distributions of altitude (<b>a</b>); slope (<b>b</b>); aspect (<b>c</b>) and LAI (<b>d</b>) at 500 m and 1 km.</p>
Full article ">Figure 4
<p>Density scatterplot between MODIS GPP and downscaled GPP at 500 m during Julian days 169–209. (<b>a</b>) represents the comparison between MODIS GPP products; and (<b>b</b>–<b>e</b>) represent the comparison between downscaled GPP and MODIS GPP at 500 m. The solid lines are the regression lines, while the dashed lines are the 1:1 lines. The green and red represent the low-density and high-density areas, respectively.</p>
Full article ">Figure 5
<p>Relationships between topographic factors and inconsistency in the MODIS GPP at 500 m and 1 km during Julian days 169–209.</p>
Full article ">Figure 6
<p>The spatial distribution of absolute GPP differences between the 500 m MODIS GPP and the downscaled GPP results for Julian days 169–209. Figures (<b>a</b>–<b>f</b>) indicate Julian days 169–209, respectively.</p>
Full article ">Figure 7
<p>The density distributions concerning the GPP differences between the 500 m MODIS GPP and the downscaled GPP results for Julian days 169–209. Figures (<b>a</b>–<b>f</b>) indicate Julian days 169–209, respectively.</p>
Full article ">
15 pages, 32418 KiB  
Article
Spatio-Temporal Analysis and Uncertainty of Fractional Vegetation Cover Change over Northern China during 2001–2012 Based on Multiple Vegetation Data Sets
by Linqing Yang, Kun Jia, Shunlin Liang, Meng Liu, Xiangqin Wei, Yunjun Yao, Xiaotong Zhang and Duanyang Liu
Remote Sens. 2018, 10(4), 549; https://doi.org/10.3390/rs10040549 - 3 Apr 2018
Cited by 34 | Viewed by 5032
Abstract
Northern China is one of the most sensitive and vulnerable regions in the country. To combat environmental degradation in northern China, a series of vegetation protection programs, such as the Three-North Shelter Forest Program (TNFSP), have been implemented. Whether the implementation of these [...] Read more.
Northern China is one of the most sensitive and vulnerable regions in the country. To combat environmental degradation in northern China, a series of vegetation protection programs, such as the Three-North Shelter Forest Program (TNFSP), have been implemented. Whether the implementation of these programs in northern China has improved the vegetation conditions has merited global attention. Therefore, quantifying vegetation changes in northern China is essential for meteorological, hydrological, ecological, and societal implications. Fractional vegetation cover (FVC) is a crucial biophysical parameter which describes land surface vegetation conditions. In this study, four FVC data sets derived from remote sensing data over northern China are employed for a spatio-temporal analysis to determine the uncertainty of fractional vegetation cover change from 2001 to 2012. Trend analysis of these data sets (including an annually varying estimate of error) reveals that FVC has increased at the rate of 0.26 ± 0.13%, 0.30 ± 0.25%, 0.12 ± 0.03%, 0.49 ± 0.21% per year in northern China, Northeast China, Northwest China, and North China during the period 2001–2012, respectively. In all of northern China, only 33.03% of pixels showed a significant increase in vegetation cover whereas approximately 16.81% of pixels showed a significant decrease and 50.16% remained relatively stable. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of the research area (The green area represents northeast China, the pink area for North China, the blue area for Northwest China, respectively).</p>
Full article ">Figure 2
<p>The spatial patterns of maximum FVC over the period 2001–2012 in northern China. (<b>a</b>) GLASS FVC; (<b>b</b>) TRAGL FVC; (<b>c</b>) GEOV1 FVC; (<b>d</b>) Li FVC.</p>
Full article ">Figure 3
<p>The temporal trends of annual maximum FVC in northern China during the periods 2001–2012. Left column is the temporal trends of (<b>a</b>) GLASS FVC, (<b>b</b>) TRAGL FVC, (<b>c</b>) GEOV1 FVC, (<b>d</b>) Li FVC. Right column is the temporal trends of annual maximum FVC that passed the significant test in northern China during the periods 2001–2012. (<b>e</b>) GLASS FVC; (<b>f</b>) TRAGL FVC; (<b>g</b>) GEOV1 FVC; (<b>h</b>) Li FVC.</p>
Full article ">Figure 4
<p>The averages of annual mean maximum FVC of four data sets over the period 2001–2012. (<b>a</b>) northern China; (<b>b</b>) Northeast China; (<b>c</b>) Northwest China; (<b>d</b>) North China.</p>
Full article ">Figure 5
<p>Correlation and root mean square error (RMSE) of each FVC data set with the averaged FVC from the three other FVC data sets used in this study. Left column is the correlation of (<b>a</b>) GLASS FVC, (<b>b</b>) TRAGL FVC, (<b>c</b>) GEOV1 FVC, (<b>d</b>) Li FVC with the average from the other FVC data sets. Right column is the corresponding RMSE value of (<b>e</b>) GLASS FVC, (<b>f</b>) TRAGL FVC, (<b>g</b>) GEOV1 FVC, (<b>h</b>) Li FVC with the average from the other FVC data sets.</p>
Full article ">Figure 6
<p>(<b>a</b>) The temporal trends of annual averaged FVC of four data sets in northern China during the period 2001–2012. (<b>b</b>) The temporal trends of annual averaged FVC of four data sets that passed the significant test in northern China during the period 2001–2012.</p>
Full article ">Figure 7
<p>Spatial pattern of mean maximum FVC (<b>a</b>) and standard error (<b>b</b>) of multi-data sets over the northern China during the period 2001–2012.</p>
Full article ">Figure 8
<p>The averages of annual mean maximum FVC and trend in FVC with the error bars showing the standard error of multi-data set average. (<b>a</b>) northern China; (<b>b</b>) Northeast China; (<b>c</b>) Northwest China; (<b>d</b>) North China.</p>
Full article ">
24 pages, 61898 KiB  
Article
Local Effects of Forests on Temperatures across Europe
by Bijian Tang, Xiang Zhao and Wenqian Zhao
Remote Sens. 2018, 10(4), 529; https://doi.org/10.3390/rs10040529 - 29 Mar 2018
Cited by 23 | Viewed by 5989
Abstract
Forests affect local climate through biophysical processes in terrestrial ecosystems. Due to the spatial and temporal heterogeneity of ecosystems in Europe, climate responses to forests vary considerably with diverse geographic and seasonal patterns. Few studies have used an empirical analysis to examine the [...] Read more.
Forests affect local climate through biophysical processes in terrestrial ecosystems. Due to the spatial and temporal heterogeneity of ecosystems in Europe, climate responses to forests vary considerably with diverse geographic and seasonal patterns. Few studies have used an empirical analysis to examine the effect of forests on temperature and the role of the background climate in Europe. In this study, we aimed to quantitatively determine the effects of forest on temperature in different seasons with MODIS (MODerate-resolution Imaging Spectroradiometer) land surface temperature (LST) data and in situ air temperature measurements. First, we compared the differences in LSTs between forests and nearby open land. Then, we paired 48 flux sites with nearby weather stations to quantify the effects of forests on surface air temperature. Finally, we explored the role of background temperatures on the above forests effects. The results showed that (1) forest in Europe generally increased LST and air temperature in northeastern Europe and decreased LST and air temperature in other areas; (2) the daytime cooling effect was dominate and produced a net cooling effect from forests in the warm season. In the cold season, daytime and nighttime warming effects drove the net effect of forests; (3) the effects of forests on temperatures were mainly negatively correlated with the background temperatures in Europe. Under extreme climate conditions, the cooling effect of forests will be stronger during heatwaves or weaker during cold spring seasons; (4) the background temperature affects the spatiotemporal distribution of differences in albedo and evapotranspiration (forest minus open land), which determines the spatial, seasonal and interannual effects of forests on temperature. The extrapolation of the results could contribute not only to model validation and development but also to appropriate land use policies for future decades under the background of global warming. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The spatial distributions of land cover types, paired sites and selected windows. The green and orange backgrounds refer to areas with forests and open lands, respectively. The paired sites are marked with red triangles. The blue points refer to the selected windows (0.4 × 0.4°), which have areas with more than 10% of forests and open land. The small panels in the below and right show the sample window numbers at each 1° (longitude and latitude) band.</p>
Full article ">Figure 2
<p>The spatial distributions of the annual mean (<b>a</b>) daytime, (<b>b</b>) nighttime and (<b>c</b>) daily average ΔLST (forest minus open land) in Europe during the period 2003–2016. The small panels in the below and right of each ΔLST show the longitudinal and latitudinal zonal average of each ΔLST for every 1° bin. The blue lines represent the 95% confidence interval (CI) estimated by the t-test.</p>
Full article ">Figure 3
<p>The spatial distributions of the annual mean (<b>a</b>) maximum, (<b>b</b>) minimum and (<b>c</b>) daily average ΔT (forest minus open land) in Europe. The small panels in the below and right show the longitudinal and latitudinal zonal average of each ΔT for every 1° bin. The background color refers to the elevation, which gradually increases from black to white.</p>
Full article ">Figure 4
<p>Spatiotemporal patterns of latitudinal variations in (<b>a</b>) daytime, (<b>b</b>) nighttime, (<b>c</b>) daily LST differences (forest minus open land) and longitudinal variations in (<b>d</b>) daytime, (<b>e</b>) nighttime and (<b>f</b>) daily LST differences (forest minus open land) during the period 2003–2016. Grids with cross symbols indicate that the LST differences are significant at the 95% CI by the <span class="html-italic">t</span>-test.</p>
Full article ">Figure 5
<p>Comparison of seasonal variations in daily maximum, daily minimum and daily mean temperature differences in three latitudinal (<b>a</b>) south of 45°N, (<b>c</b>) between 45°N and 55°N, (<b>e</b>) north of 55°N, and longitudinal (<b>b</b>) west of 5°E, (<b>d</b>) between 5°E and 15°E, (<b>f</b>) east of 15°E ranges. The red solid and red dashed lines indicate Tmax for forests and open lands, respectively. The blue solid and blue dashed lines indicate Tmin for forests and open lands, respectively. The black solid line indicates the Tmean difference from forests minus open lands.</p>
Full article ">Figure 6
<p>The relationships between (<b>a</b>) background LST and daily LST differences (forest minus open land) and (<b>b</b>) background air temperature and daily mean air temperature differences (forest minus open land). The daily LST differences are binned and averaged on 1° background LST intervals (i.e., the LST for all pixels within a window). The daily mean air temperature differences are binned and averaged on 1° grids for background air temperature (i.e., air temperatures of forest sites). The thin black bars represent the 95% confidence interval (CI) by the <span class="html-italic">t</span>-test.</p>
Full article ">Figure 7
<p>The relationship between background LST and daily LST differences (forest minus open land) during various years. (<b>a</b>) The rate of change for daily LST differences under different background LSTs. The number 0.1 indicates that the daily LST difference increases by 0.1 °C when the background LST increases by 1 °C. (<b>b</b>) Significance of the relationship between background LST and daily LST differences.</p>
Full article ">Figure 8
<p>The relationship between background LST and daily LST differences (forest minus open land) in (<b>a</b>) spring, (<b>b</b>) summer, (<b>c</b>) autumn, and (<b>d</b>) winter during various years. Spring, summer, autumn and winter are defined by March and May, June and August, September and November, and December and February, respectively. The significance map of the four seasons is similar to <a href="#remotesensing-10-00529-f007" class="html-fig">Figure 7</a>b and is not shown here.</p>
Full article ">Figure 9
<p>The spatial distributions of annual mean (<b>a</b>) albedo (%) and (<b>b</b>) ET differences (forest minus open land) in Europe. The periods used to analyze albedo and ET differences are 2003–2016 and 2003–2014, respectively. The below and right small panels for each difference show the longitudinal and latitudinal zonal averages for every 1° bin. The blue lines represent the 95% confidence interval (CI) estimated by the <span class="html-italic">t</span>-test.</p>
Full article ">Figure 10
<p>Spatiotemporal patterns of (<b>a</b>) latitudinal and (<b>c</b>) longitudinal variations in albedo (%) differences (forest minus open land) during the period 2003–2016 and (<b>b</b>) latitudinal and (<b>d</b>) longitudinal variations in ET differences (forest minus open land) during the period 2003–2014. Grids with cross symbols indicate differences that are significant at the 95% CI by the <span class="html-italic">t</span>-test.</p>
Full article ">Figure 11
<p>The relationship between the annual mean background LST and the annual mean (<b>a</b>) albedo (%) differences (forest minus open land) and (<b>b</b>) ET differences (forest minus open land) during various years. The number 0.1 in (<b>a</b>,<b>b</b>) indicates that the albedo and ET differences increase by 0.1% and 0.1 mm/day, respectively, when the background LST increases by 1 °C. Only when the valid year of a window was greater than five years was the relationship calculated.</p>
Full article ">Figure 11 Cont.
<p>The relationship between the annual mean background LST and the annual mean (<b>a</b>) albedo (%) differences (forest minus open land) and (<b>b</b>) ET differences (forest minus open land) during various years. The number 0.1 in (<b>a</b>,<b>b</b>) indicates that the albedo and ET differences increase by 0.1% and 0.1 mm/day, respectively, when the background LST increases by 1 °C. Only when the valid year of a window was greater than five years was the relationship calculated.</p>
Full article ">Figure 12
<p>Annual mean air temperatures at different latitudes in the Northern Hemisphere (north of 10°N) during the period 2003–2016, where NA represents North America, EU represents Europe, AS represents Asia and NH represents the Northern Hemisphere. The dotted line represents 6.5 °C.</p>
Full article ">Figure 13
<p>Interannual variability of (<b>a</b>) the seasonal mean LST in summer, the difference in ET (forest minus open land) in summer, and background LST in summer and (<b>b</b>) the seasonal mean LST in spring, the difference in albedo (forest minus open land) in spring, and the background LST in spring.</p>
Full article ">
22 pages, 64129 KiB  
Article
Assessment of Methods for Passive Microwave Snow Cover Mapping Using FY-3C/MWRI Data in China
by Xiaojing Liu, Lingmei Jiang, Shengli Wu, Shirui Hao, Gongxue Wang and Jianwei Yang
Remote Sens. 2018, 10(4), 524; https://doi.org/10.3390/rs10040524 - 27 Mar 2018
Cited by 17 | Viewed by 4361
Abstract
Ongoing information on snow and its extent is critical for understanding global water and energy cycles. Passive microwave data have been widely used in snow cover mapping given their long-time observation capabilities under all-weather conditions. However, assessments of different passive microwave (PMW) snow [...] Read more.
Ongoing information on snow and its extent is critical for understanding global water and energy cycles. Passive microwave data have been widely used in snow cover mapping given their long-time observation capabilities under all-weather conditions. However, assessments of different passive microwave (PMW) snow cover area (SCA) mapping algorithms have rarely been reported, especially in China. In this study, the performances of seven PMW SCA mapping algorithms were tested using in situ snow depth measurements and a one-kilometer Interactive Multisensor Snow and Ice Mapping System (IMS) snow cover product over China. The selected algorithms are the FY3 algorithm, Grody’s algorithm, the South China algorithm, Kelly’s algorithm, Singh’s algorithm, Hall’s algorithm and Neal’s algorithm. During the test period, most algorithms performed reasonably well. The overall accuracy of all algorithms is higher than 0.895 against in situ observations and higher than 0.713 against the IMS product. In general, Singh’s algorithm, Hall’s algorithm and Neal’s algorithm had poor performance during the test. Their misclassification errors were larger than those of the remaining algorithms. Grody’s algorithm, the South China algorithm and Kelly’s algorithm had higher positive predictive values and lower omission errors than those of the others. The errors of these three algorithms were mainly caused by variations in commission errors. Comparing to Grody’s algorithm, the South China algorithm and Kelly’s algorithm, the FY3 algorithm presented a conservative snow cover estimation to balance the problem between snow identification and overestimation. As a result, the overall accuracy of the FY3 algorithm was the highest of all the tested algorithms. The accuracy of all algorithms tended to decline with a decreased snow cover fraction as well as SD < 5 cm. All tested algorithms have severe omission errors over barren land and grasslands. The results shown in this study contribute to ongoing efforts to improve the performance and applicability of PMW SCA algorithms. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chinese meteorological stations used in this work.</p>
Full article ">Figure 2
<p>Comparison of PMW SCA maps for 7 January 2014: (<b>a</b>) FY SCA map, (<b>b</b>) Grody’s SCA map, (<b>c</b>) Hall’s SCA map, (<b>d</b>) Kelly’s SCA map, (<b>e</b>) Neal’s SCA map, (<b>f</b>) Singh’s SCA map, (<b>g</b>) the South China SCA map. (blended SCA image (<b>g</b>) using both ascending and descending data for the South China algorithm and descending SCA images(<b>a</b>–<b>f</b>) for the remaining PMW SCA mapping algorithms).</p>
Full article ">Figure 3
<p>Monthly OA (<b>a</b>), PPV (<b>b</b>), OE (<b>c</b>) and CE (<b>d</b>) of the seven PMW SCA maps based on in situ SD observations.</p>
Full article ">Figure 4
<p>Monthly OA (<b>a</b>), PPV (<b>b</b>), OE (<b>c</b>) and CE (<b>d</b>) of the seven PMW SCA maps based on the IMS snow cover product.</p>
Full article ">Figure 5
<p>Globeland30 land-cover map.</p>
Full article ">Figure 6
<p>OA (<b>a</b>), PPV (<b>b</b>), OE (<b>c</b>) and CE (<b>d</b>) of the seven PMW SCA maps compared to station observations in different land-cover types.</p>
Full article ">Figure 7
<p>OE (<b>a</b>) and PPV (<b>b</b>) of the PMW SCA mapping algorithms associated with different SCFs.</p>
Full article ">Figure 8
<p>OE (<b>a</b>) and PPV (<b>b</b>) of the seven PMW SCA maps compared to ground observations at different snow depths.</p>
Full article ">Figure 9
<p>The evaluation results of classification criteria for each algorithm: PPV (<b>a</b>), CE (<b>b</b>), OE (<b>c</b>).</p>
Full article ">Figure 9 Cont.
<p>The evaluation results of classification criteria for each algorithm: PPV (<b>a</b>), CE (<b>b</b>), OE (<b>c</b>).</p>
Full article ">Figure 10
<p>Box plots of the single-band indexes (<b>a1</b>,<b>a2</b>), the Tb gradient indexes (<b>b1</b>,<b>b2</b>), the polarization difference indexes (<b>c1</b>,<b>c2</b>), the polarization ratio indexes(<b>d1</b>,<b>d2</b>) and the day–night Tb difference indexes (<b>e</b>).</p>
Full article ">Figure 11
<p>Box plots of Tb index values and thresholds of the seven PMW SCA mapping algorithms: Grody’s algorithm (<b>a1</b>,<b>a2</b>), Kelly’s SCA algorithm (<b>b1</b>,<b>b2</b>), the South China algorithm (<b>c1</b>–<b>c3</b>), FY SCA algorithm (<b>d1</b>,<b>d2</b>), Neal’s algorithm (<b>e1</b>,<b>e2</b>), Singh’s algorithm (<b>f1</b>,<b>f2</b>), Hall’s algorithm (<b>g1</b>,<b>g2</b>). (The up and down arrows indicate the lower bounds and upper bounds of the thresholds, respectively. Arrows in red and in blue indicate thresholds for snow identification and non-snow identification, respectively.)</p>
Full article ">Figure 11 Cont.
<p>Box plots of Tb index values and thresholds of the seven PMW SCA mapping algorithms: Grody’s algorithm (<b>a1</b>,<b>a2</b>), Kelly’s SCA algorithm (<b>b1</b>,<b>b2</b>), the South China algorithm (<b>c1</b>–<b>c3</b>), FY SCA algorithm (<b>d1</b>,<b>d2</b>), Neal’s algorithm (<b>e1</b>,<b>e2</b>), Singh’s algorithm (<b>f1</b>,<b>f2</b>), Hall’s algorithm (<b>g1</b>,<b>g2</b>). (The up and down arrows indicate the lower bounds and upper bounds of the thresholds, respectively. Arrows in red and in blue indicate thresholds for snow identification and non-snow identification, respectively.)</p>
Full article ">
17 pages, 25990 KiB  
Article
Individual and Interactive Influences of Anthropogenic and Ecological Factors on Forest PM2.5 Concentrations at an Urban Scale
by Guoliang Yun, Shudi Zuo, Shaoqing Dai, Xiaodong Song, Chengdong Xu, Yilan Liao, Peiqiang Zhao, Weiyin Chang, Qi Chen, Yaying Li, Jianfeng Tang, Wang Man and Yin Ren
Remote Sens. 2018, 10(4), 521; https://doi.org/10.3390/rs10040521 - 26 Mar 2018
Cited by 21 | Viewed by 5050
Abstract
Integration of Landsat images and multisource data using spatial statistical analysis and geographical detector models can reveal the individual and interactive influences of anthropogenic activities and ecological factors on concentrations of atmospheric particulate matter less than 2.5 microns in diameter (PM2.5). [...] Read more.
Integration of Landsat images and multisource data using spatial statistical analysis and geographical detector models can reveal the individual and interactive influences of anthropogenic activities and ecological factors on concentrations of atmospheric particulate matter less than 2.5 microns in diameter (PM2.5). This approach has been used in many studies to estimate biomass and forest disturbance patterns and to monitor carbon sinks. However, the approach has rarely been used to comprehensively analyze the individual and interactive influences of anthropogenic factors (e.g., population density, impervious surface percentage) and ecological factors (e.g., canopy density, stand age, and elevation) on PM2.5 concentrations. To do this, we used Landsat-8 images and meteorological data to retrieve quantitative data on the concentrations of particulates (PM2.5), then integrated a forest management planning inventory (FMPI), population density distribution data, meteorological data, and topographic data in a Geographic Information System database, and applied a spatial statistical analysis model to identify aggregated areas (hot spots and cold spots) of particulates in the urban area of Jinjiang city, China. A geographical detector model was used to analyze the individual and interactive influences of anthropogenic and ecological factors on PM2.5 concentrations. We found that particulate concentration hot spots are mainly distributed in urban centers and suburbs, while cold spots are mainly distributed in the suburbs and exurban region. Elevation was the dominant individual factor affecting PM2.5 concentrations, followed by dominant tree species and meteorological factors. A combination of human activities (e.g., population density, impervious surface percentage) and multiple ecological factors caused the dominant interactive effects, resulting in increased PM2.5 concentrations. Our study suggests that human activities and multiple ecological factors effect PM2.5 concentrations both individually and interactively. We conclude that in order to reveal the direct and indirect effects of human activities and multiple factors on PM2.5 concentrations in urban forests, quantification of fusion satellite data and spatial statistical methods should be conducted in urban areas. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Correlation between <math display="inline"> <semantics> <mrow> <msub> <mi>k</mi> <mrow> <mi>A</mi> <mi>O</mi> <mi>D</mi> <mo>,</mo> <mi>D</mi> <mi>r</mi> <mi>y</mi> </mrow> </msub> </mrow> </semantics> </math> and PM<sub>2.5</sub> concentrations on 13 December 2014, 29 December 2014, and 14 January 2015 (both acquired from Environmental Protection of Jinjiang <math display="inline"> <semantics> <mrow> <msub> <mi>k</mi> <mrow> <mi>A</mi> <mi>O</mi> <mi>D</mi> <mo>,</mo> <mi>D</mi> <mi>r</mi> <mi>y</mi> </mrow> </msub> </mrow> </semantics> </math> represents the aerosol extinction coefficient in dry conditions).</p>
Full article ">Figure 2
<p>Spatial distributions of urban forest PM<sub>2.5</sub> concentrations on three different days at the optimal distance threshold of 3500 m.</p>
Full article ">Figure 3
<p>The spatial distribution of population density in 2014 (mapped using kernel density in Arc GIS 10.3).</p>
Full article ">Figure 4
<p>The power of determinants of different impact factors (forest, soils, topography, meteorological factors, and population) on PM<sub>2.5</sub> concentrations in Jinjiang. (PA = patch area, DS = dominant species, CD = canopy density, SA = stand age, SI = site index, SD = soil depth, HD = humus depth, ELE = elevation, SDe = degree of slope, SPo = slope of position, SDi = slope aspect, PopD = population density, ISP = impervious surface percentage, TEM = temperature, SH = specific humidity, PS = pressure, WS = wind speed), <span class="html-italic">p</span>-values &lt; 0.1 for all factors.</p>
Full article ">Figure 5
<p>Scores for the interactive effects of impact factors sensitivity rankings at the three periods studied calculated based on simulations using multisource data ((<b>A</b>) 13 December 2014; (<b>B</b>) 29 December 2014; (<b>C</b>) 14 January 2015).</p>
Full article ">
22 pages, 38594 KiB  
Article
Spatiotemporal Dynamics in Vegetation GPP over the Great Khingan Mountains Using GLASS Products from 1982 to 2015
by Ling Hu, Wenjie Fan, Huazhong Ren, Suhong Liu, Yaokui Cui and Peng Zhao
Remote Sens. 2018, 10(3), 488; https://doi.org/10.3390/rs10030488 - 20 Mar 2018
Cited by 39 | Viewed by 6446
Abstract
Gross primary productivity (GPP) is an important parameter that represents the productivity of vegetation and responses to various ecological environments. The Greater Khingan Mountain (GKM) is one of the most important state-owned forest bases, and boreal forests, including the largest primeval cold-temperature bright [...] Read more.
Gross primary productivity (GPP) is an important parameter that represents the productivity of vegetation and responses to various ecological environments. The Greater Khingan Mountain (GKM) is one of the most important state-owned forest bases, and boreal forests, including the largest primeval cold-temperature bright coniferous forest in China, are widely distributed in the GKM. This study aimed to reveal spatiotemporal vegetation variations in the GKM on the basis of GPP products that were generated by the Global LAnd Surface Satellite (GLASS) program from 1982 to 2015. First, we explored the spatiotemporal distribution of vegetation across the GKM. Then we analyzed the relationships between GPP variation and driving factors, including meteorological elements, growing season length (GSL), and Fraction of Photosynthetically Active Radiation (FPAR), to investigate the dominant factor for GPP dynamics. Results demonstrated that (1) the spatial distribution of accumulated GPP (AG) in spring, summer, autumn, and the growing season varied due to three main reasons: understory vegetation, altitude, and land cover; (2) interannual AG in summer, autumn, and the growing season significantly increased at the regional scale during the past 34 years under climate warming and drying; (3) interannual changes of accumulated GPP in the growing season (AGG) at the pixel scale displayed a rapid expansion in areas with a significant increasing trend (p < 0.05) during the period of 1982–2015 and this trend was caused by the natural forest protection project launched in 1998; and finally, (4) an analysis of driving factors showed that daily sunshine duration in summer was the most important factor for GPP in the GKM and this is different from previous studies, which reported that the GSL plays a crucial role in other areas. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Distribution of vegetation type and location of the Greater Khingan Mountains (GKM) in China, based on the Vegetation Regionalization Map of China (1:6,000,000) [<a href="#B47-remotesensing-10-00488" class="html-bibr">47</a>]; (<b>b</b>) DEM(Digital Elevation Model) and the meteorological station of the GKM; and, (<b>c</b>) Land cover of the GKM. The land cover data is from the MODIS dataset of 2001.</p>
Full article ">Figure 2
<p>Spatial distribution of (<b>a</b>) spring-average accumulated gross primary productivity (AG); (<b>b</b>) summer-average AG; (<b>c</b>) autumn-average AG; and, (<b>d</b>) growing season (GS)-average AG during 1982–2015.</p>
Full article ">Figure 3
<p>Intra-annual variation characteristic of gross primary productivity (GPP) values during 1982–2015.</p>
Full article ">Figure 4
<p>Interannual variation trends of AG in the GKM from 1982 to 2015.</p>
Full article ">Figure 4 Cont.
<p>Interannual variation trends of AG in the GKM from 1982 to 2015.</p>
Full article ">Figure 5
<p>Area fraction of the AGG displaying different statistical significance levels with progressively longer time series since 1982 (%). Significant positive stands for the increasing of GPP and significant negative refers to the reduction of GPP; both indicate a statistical significance of the linear regression with a Pearson correlation less than 0.05.</p>
Full article ">Figure 6
<p>Spatial distribution of AGG trends across the GKM during the periods of (<b>a</b>) 1982–1999; (<b>b</b>) 1982–2005; (<b>c</b>) 1982–2010; and, (<b>d</b>) 1982–2015. The linear trends of the AGG were calculated at the 95% confidence level.</p>
Full article ">Figure 7
<p>Variation in phenological parameters in the GKM from 1982 to 2014 (SOG means starting day of growing season, EOG means ending day of growing season, and GSL means growing season length).</p>
Full article ">Figure 8
<p>Correlation coefficient between GSL and AGG in the GKM during 1982–2014. (Green indicates significant positive correlation (r &gt; 0, <span class="html-italic">p</span> &lt; 0.05), brown indicates significant negative correlation (r &lt; 0, <span class="html-italic">p</span> &lt; 0.05), and gray indicates non-significant correlation (<span class="html-italic">p</span> &gt; 0.05)).</p>
Full article ">Figure 9
<p>Correlation coefficient between Fraction of Photosynthetically Active Radiation (FPAR) and AGG in the GKM during 1982–2014. (Yellow and red indicate significant positive correlation (r &gt; 0, <span class="html-italic">p</span> &lt; 0.05), blue indicates significant negative correlation (r &lt; 0, <span class="html-italic">p</span> &lt; 0.05), and gray indicates non-significant correlation (<span class="html-italic">p</span> &gt; 0.05)).</p>
Full article ">Figure 10
<p>Spatial distribution of (<b>a</b>) dominant and (<b>b</b>) subdominant factors for AGG in the GKM from 1982 to 2015 (Non-significant means that there is no significantly effective factor for AGG among these five driving factors according to the <span class="html-italic">F</span>-test and <span class="html-italic">t</span>-test. A significant <span class="html-italic">F</span>-test was under <span class="html-italic">p</span> &lt; 0.05, and a significant <span class="html-italic">t</span>-test was under the absolute value of <span class="html-italic">t</span> ≥ 1.96).</p>
Full article ">Figure 11
<p>Contrast of AG in specific land cover and other areas in the GKM.</p>
Full article ">Figure 12
<p>Variation of AGG with elevation.</p>
Full article ">Figure 13
<p>Variation in annual mean daily temperature and annual mean daily precipitation in the growing season over the GKM from 1982 to 2015.</p>
Full article ">Figure 14
<p>Variation in timber yield in Heilongjiang Province since 1980.</p>
Full article ">Figure 15
<p>Burnt zone of the fire in 1987 and land cover in the GKM.</p>
Full article ">
17 pages, 13949 KiB  
Article
Estimation of Daily Average Downward Shortwave Radiation over Antarctica
by Yingji Zhou, Guangjian Yan, Jing Zhao, Qing Chu, Yanan Liu, Kai Yan, Yiyi Tong, Xihan Mu, Donghui Xie and Wuming Zhang
Remote Sens. 2018, 10(3), 422; https://doi.org/10.3390/rs10030422 - 9 Mar 2018
Cited by 12 | Viewed by 6173
Abstract
Surface shortwave (SW) irradiation is the primary driving force of energy exchange in the atmosphere and land interface. The global climate is profoundly influenced by irradiation changes due to the special climatic condition in Antarctica. Remote-sensing retrieval can offer only the instantaneous values [...] Read more.
Surface shortwave (SW) irradiation is the primary driving force of energy exchange in the atmosphere and land interface. The global climate is profoundly influenced by irradiation changes due to the special climatic condition in Antarctica. Remote-sensing retrieval can offer only the instantaneous values in an area, whilst daily cycle and average values are necessary for further studies and applications, including climate change, ecology, and land surface process. When considering the large values of and small diurnal changes of solar zenith angle and cloud coverage, we develop two methods for the temporal extension of remotely sensed downward SW irradiance over Antarctica. The first one is an improved sinusoidal method, and the second one is an interpolation method based on cloud fraction change. The instantaneous irradiance data and cloud products are used in both methods to extend the diurnal cycle, and obtain the daily average value. Data from South Pole and Georg von Neumayer stations are used to validate the estimated value. The coefficient of determination (R2) between the estimated daily averages and the measured values based on the first method is 0.93, and the root mean square error (RMSE) is 32.21 W/m2 (8.52%). As for the traditional sinusoidal method, the R2 and RMSE are 0.68 and 70.32 W/m2 (18.59%), respectively The R2 and RMSE of the second method are 0.96 and 25.27 W/m2 (6.98%), respectively. These values are better than those of the traditional linear interpolation (0.79 and 57.40 W/m2 (15.87%)). Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Overpass times of the Suomi National Polar-Orbiting Partnership satellite in one day and distribution of Baseline Surface Radiation Network stations in Antarctica.</p>
Full article ">Figure 2
<p>(<b>a</b>) Diurnal variation range of solar zenith angle (SZA) and downward surface shortwave irradiance (DSSR) at the Georg von Neumayer (GVN) station on 17 October 2013; (<b>b</b>) Diurnal variation range of SZA and DSSR at the GVN station on 22 December 2013; (<b>c</b>) Diurnal variation range of SZA and DSSR at the South Pole (SPO) station on 17 October 2013; and, (<b>d</b>) Diurnal variation range of SZA and DSSR at SPO station on 22 December 2013.</p>
Full article ">Figure 3
<p>Process of calculating daily average downward shortwave irradiation values.</p>
Full article ">Figure 4
<p>Change range of solar zenith angle at latitudes of 60°S, 70°S, 80°S, and 90°S in all year.</p>
Full article ">Figure 5
<p>Results of hourly interpolation in the Georg von Neumayer station area on 9 December 2014.</p>
Full article ">Figure 6
<p>Results of hourly interpolation in the South Pole station area on 2 January 2015.</p>
Full article ">Figure 7
<p>(<b>a</b>) Daily average value of solar global irradiation over horizontal surface in the Georg von Neumayer station area on 9 December 2014; and, (<b>b</b>) Daily average value of solar global irradiation over horizontal surface in the South Pole station area on 2 January 2015.</p>
Full article ">Figure 8
<p>(<b>a</b>) Comparison of the estimated downward shortwave irradiation from the improved sinusoidal method and the ground-measured downward shortwave irradiation in the Georg von Neumayer station; and, (<b>b</b>) Comparison of the estimated downward shortwave irradiation from the cloud coverage fraction interpolated method and the ground-measured downward shortwave irradiation in the South Pole station area.</p>
Full article ">Figure A1
<p>Cloud pixel and target pixel in the slide window.</p>
Full article ">Figure A2
<p>The Classification for sun-sensor-hemisphere cloud cover conditions.</p>
Full article ">
19 pages, 2257 KiB  
Article
A Lookup-Table-Based Approach to Estimating Surface Solar Irradiance from Geostationary and Polar-Orbiting Satellite Data
by Hailong Zhang, Chong Huang, Shanshan Yu, Li Li, Xiaozhou Xin and Qinhuo Liu
Remote Sens. 2018, 10(3), 411; https://doi.org/10.3390/rs10030411 - 7 Mar 2018
Cited by 15 | Viewed by 8050
Abstract
Incoming surface solar irradiance (SSI) is essential for calculating Earth’s surface radiation budget and is a key parameter for terrestrial ecological modeling and climate change research. Remote sensing images from geostationary and polar-orbiting satellites provide an opportunity for SSI estimation through directly retrieving [...] Read more.
Incoming surface solar irradiance (SSI) is essential for calculating Earth’s surface radiation budget and is a key parameter for terrestrial ecological modeling and climate change research. Remote sensing images from geostationary and polar-orbiting satellites provide an opportunity for SSI estimation through directly retrieving atmospheric and land-surface parameters. This paper presents a new scheme for estimating SSI from the visible and infrared channels of geostationary meteorological and polar-orbiting satellite data. Aerosol optical thickness and cloud microphysical parameters were retrieved from Geostationary Operational Environmental Satellite (GOES) system images by interpolating lookup tables of clear and cloudy skies, respectively. SSI was estimated using pre-calculated offline lookup tables with different atmospheric input data of clear and cloudy skies. The lookup tables were created via the comprehensive radiative transfer model, Santa Barbara Discrete Ordinate Radiative Transfer (SBDART), to balance computational efficiency and accuracy. The atmospheric attenuation effects considered in our approach were water vapor absorption and aerosol extinction for clear skies, while cloud parameters were the only atmospheric input for cloudy-sky SSI estimation. The approach was validated using one-year pyranometer measurements from seven stations in the SURFRAD (SURFace RADiation budget network). The results of the comparison for 2012 showed that the estimated SSI agreed with ground measurements with correlation coefficients of 0.94, 0.69, and 0.89 with a bias of 26.4 W/m2, −5.9 W/m2, and 14.9 W/m2 for clear-sky, cloudy-sky, and all-sky conditions, respectively. The overall root mean square error (RMSE) of instantaneous SSI was 80.0 W/m2 (16.8%), 127.6 W/m2 (55.1%), and 99.5 W/m2 (25.5%) for clear-sky, cloudy-sky (overcast sky and partly cloudy sky), and all-sky (clear-sky and cloudy-sky) conditions, respectively. A comparison with other state-of-the-art studies suggests that our proposed method can successfully estimate SSI with a maximum improvement of an RMSE of 24 W/m2. The clear-sky SSI retrieval was sensitive to aerosol optical thickness, which was largely dependent on the diurnal surface reflectance accuracy. Uncertainty in the pre-defined horizontal visibility for ‘clearest sky’ will eventually lead to considerable SSI retrieval error. Compared to cloud effective radius, the retrieval error of cloud optical thickness was a primary factor that determined the SSI estimation accuracy for cloudy skies. Our proposed method can be used to estimate SSI for clear and one-layer cloud sky, but is not suitable for multi-layer clouds overlap conditions as a lower-level cloud cannot be detected by the optical sensor when a higher-level cloud has a higher optical thickness. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Estimated sensor-received radiance for multi-layer clouds overlap conditions.</p>
Full article ">Figure 2
<p>Estimated surface-received radiance for multi-layer clouds overlap conditions.</p>
Full article ">Figure 3
<p>Flow chart of surface solar irradiance (SSI) retrieval from geostationary and polar-orbiting satellite data. MODIS: Moderate Resolution Imaging Spectroradiometer; GOES: Geostationary Operational Environmental Satellite; NCEP CFSR: National Centers for Environmental Prediction Climate Forecast System Reanalysis; LUT: lookup table; AOD: aerosol optical density; DEM: digital elevation model.</p>
Full article ">Figure 4
<p>Validation results for the instantaneous surface solar irradiance estimated at seven Surface Radiation Budget Network (SURFRAD) stations by the scheme proposed in this study. BON: Bondville, Illinois; DRA: Desert Rock, Nevada; FPK: Fort Peck, Montana; GWN: Goodwin Creek, Mississippi; PSU: Penn State, Pennsylvania; SXF: Sioux Falls, South Dakota; TBL: Table Mountain, Colorado.</p>
Full article ">Figure 5
<p>Statistical results of observed AOD at six SURFRAD stations. The dashed lines on the vertical axis are AOD values of 0.06 for pre-defined clearest-sky conditions of diurnal surface reflectance retrieval.</p>
Full article ">Figure 6
<p>The sensitivity of SSI to cloud optical thickness (given an effective particle radius of 20 μm) and effective particle radius (given a cloud optical thickness of 20) for water clouds (<b>left</b>) and ice clouds (<b>right</b>).</p>
Full article ">Figure 7
<p>Validation of SSI for water cloud cases at SURFRAD sites (mixed and undetected clouds were not included in the comparison).</p>
Full article ">Figure 8
<p>Validation of SSI for ice cloud cases at SURFRAD sites (mixed and undetected clouds were not included in the comparison).</p>
Full article ">
22 pages, 23300 KiB  
Article
Estimation of Forest Canopy Height and Aboveground Biomass from Spaceborne LiDAR and Landsat Imageries in Maryland
by Mengjia Wang, Rui Sun and Zhiqiang Xiao
Remote Sens. 2018, 10(2), 344; https://doi.org/10.3390/rs10020344 - 23 Feb 2018
Cited by 45 | Viewed by 6793
Abstract
Mapping the regional distribution of forest canopy height and aboveground biomass is worthwhile and necessary for estimating the carbon stocks on Earth and assessing the terrestrial carbon flux. In this study, we produced maps of forest canopy height and the aboveground biomass at [...] Read more.
Mapping the regional distribution of forest canopy height and aboveground biomass is worthwhile and necessary for estimating the carbon stocks on Earth and assessing the terrestrial carbon flux. In this study, we produced maps of forest canopy height and the aboveground biomass at a 30 m spatial resolution in Maryland by combining Geoscience Laser Altimeter System (GLAS) data and Landsat spectral imageries. The processes for calculating the forest biomass included the following: (i) processing the GLAS waveform and calculating spatially discrete forest canopy heights; (ii) developing canopy height models from Landsat imagery and extrapolating them to spatially contiguous canopy heights in Maryland; and, (iii) estimating forest aboveground biomass according to the relationship between canopy height and biomass. In our study, we explore the ability to use the GLAS waveform to calculate canopy height without ground-measured forest metrics (R2 = 0.669, RMSE = 4.82 m, MRE = 15.4%). The machine learning models performed better than the principal component model when mapping the regional forest canopy height and aboveground biomass. The total forest aboveground biomass in Maryland reached approximately 160 Tg. When compared with the existing Biomass_CMS map, our biomass estimates presented a similar distribution where higher values were in the Western Shore Uplands region and Folded Application Mountain section, while lower values were located in the Delmarva Peninsula and Allegheny Mountain regions. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Overall introduction to the study area. (<b>a</b>) The elevation map and the distribution of the physical regions. (<b>b</b>) The distribution of the forest in Maryland and Geoscience Laser Altimeter System (GLAS) footprints located in forest areas.</p>
Full article ">Figure 2
<p>A typical waveform profile of a GLAS shot in Maryland.</p>
Full article ">Figure 3
<p>The evaluation of estimated canopy height by the GLAS waveform. (<b>a</b>) The evaluation of canopy height without slope correction; and, (<b>b</b>) The evaluation of canopy height with slope correction.</p>
Full article ">Figure 4
<p>PCA analysis results. (<b>a</b>) Power model; and, (<b>b</b>) Evaluation of this model.</p>
Full article ">Figure 5
<p>The evaluation results of the machine learning models. (<b>a</b>) Evaluation result of the BPANN model; (<b>b</b>) evaluation result of the SVR model; and, (<b>c</b>) evaluation result of the RF power model.</p>
Full article ">Figure 6
<p>The distribution of forest canopy height in Maryland. (<b>a</b>) Forest canopy height estimated by the first principal component power model; (<b>b</b>) Forest canopy height estimated by the BPANN model; (<b>c</b>) Forest canopy height estimated by the SVR model; and (<b>d</b>) Forest canopy height estimated by the RF model.</p>
Full article ">Figure 6 Cont.
<p>The distribution of forest canopy height in Maryland. (<b>a</b>) Forest canopy height estimated by the first principal component power model; (<b>b</b>) Forest canopy height estimated by the BPANN model; (<b>c</b>) Forest canopy height estimated by the SVR model; and (<b>d</b>) Forest canopy height estimated by the RF model.</p>
Full article ">Figure 7
<p>Forest aboveground biomass model and the evaluation results. (<b>a</b>) Power model to estimate forest aboveground biomass; and, (<b>b</b>) Evaluation result of the biomass estimation model.</p>
Full article ">Figure 8
<p>The distribution of forest aboveground biomass in Maryland. (<b>a</b>) Forest aboveground biomass estimated by the PCA power model; (<b>b</b>) Forest aboveground biomass estimated by the BP-ANN model; (<b>c</b>) Forest aboveground biomass estimated by the SVR model; (<b>d</b>) Forest aboveground biomass estimated by the RF model; and, (<b>e</b>) Forest aboveground biomass estimated by the CMS.</p>
Full article ">Figure 8 Cont.
<p>The distribution of forest aboveground biomass in Maryland. (<b>a</b>) Forest aboveground biomass estimated by the PCA power model; (<b>b</b>) Forest aboveground biomass estimated by the BP-ANN model; (<b>c</b>) Forest aboveground biomass estimated by the SVR model; (<b>d</b>) Forest aboveground biomass estimated by the RF model; and, (<b>e</b>) Forest aboveground biomass estimated by the CMS.</p>
Full article ">Figure 9
<p>The results of forest biomass difference. (<b>a</b>) The map of biomass difference in Maryland; (<b>b</b>) The statistical result of biomass difference.</p>
Full article ">Figure 9 Cont.
<p>The results of forest biomass difference. (<b>a</b>) The map of biomass difference in Maryland; (<b>b</b>) The statistical result of biomass difference.</p>
Full article ">Figure 10
<p>The forest aboveground biomass in Maryland. (<b>a</b>) Statistical forest biomass values of each county; (<b>b</b>) statistical forest biomass values of each physical region; and (<b>c</b>) the total biomass estimated by all models.</p>
Full article ">Figure 10 Cont.
<p>The forest aboveground biomass in Maryland. (<b>a</b>) Statistical forest biomass values of each county; (<b>b</b>) statistical forest biomass values of each physical region; and (<b>c</b>) the total biomass estimated by all models.</p>
Full article ">
17 pages, 2935 KiB  
Article
Monitoring Rice Phenology Based on Backscattering Characteristics of Multi-Temporal RADARSAT-2 Datasets
by Ze He, Shihua Li, Yong Wang, Leiyu Dai and Sen Lin
Remote Sens. 2018, 10(2), 340; https://doi.org/10.3390/rs10020340 - 23 Feb 2018
Cited by 57 | Viewed by 7209
Abstract
Accurate estimation and monitoring of rice phenology is necessary for the management and yield prediction of rice. The radar backscattering coefficient, one of the most direct and accessible parameters has been proved to be capable of retrieving rice growth parameters. This paper aims [...] Read more.
Accurate estimation and monitoring of rice phenology is necessary for the management and yield prediction of rice. The radar backscattering coefficient, one of the most direct and accessible parameters has been proved to be capable of retrieving rice growth parameters. This paper aims to investigate the possibility of monitoring the rice phenology (i.e., transplanting, vegetative, reproductive, and maturity) using the backscattering coefficients or their simple combinations of multi-temporal RADARSAT-2 datasets only. Four RADARSAT-2 datasets were analyzed at 30 sample plots in Meishan City, Sichuan Province, China. By exploiting the relationships of the backscattering coefficients and their combinations versus the phenology of rice, HH/VV, VV/VH, and HH/VH ratios were found to have the greatest potential for phenology monitoring. A decision tree classifier was applied to distinguish the four phenological phases, and the classifier was effective. The validation of the classifier indicated an overall accuracy level of 86.2%. Most of the errors occurred in the vegetative and reproductive phases. The corresponding errors were 21.4% and 16.7%, respectively. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The study area imaged by the French SPOT-6 optical sensor on 15 July 2016. Thirty sample sites are located and numbered.</p>
Full article ">Figure 2
<p>Four rice phenological phases in the study area in 2016. (<b>a</b>) transplanting, (<b>b</b>) vegetative, (<b>c</b>) reproductive, and (<b>d</b>) maturity.</p>
Full article ">Figure 3
<p>In situ rice phenology at 30 sample sites in 2016. Blue, green, yellow, and red dots, respectively, represent the transplanting, vegetative, reproductive, and maturity phase. Four horizontal lines denote the acquisition dates of four RADARSAT-2 datasets.</p>
Full article ">Figure 4
<p>A RADARSAT-2 image acquired on 2 July 2016. VH, VV, and HH bands were assigned as red, green, and blue colors, respectively. Study area is within the yellow box.</p>
Full article ">Figure 5
<p>A simple decision tree of a variable, λ. Three thresholds divide λ into four classes through two decision layers.</p>
Full article ">Figure 6
<p>Distributions of development phases vs. VH backscattering coefficients. The data are training data. At about −20 dB, the transplanting phase is separated from other three phases.</p>
Full article ">Figure 7
<p>Boxplots of backscattering coefficients (training dataset, expressed in logarithmic scale) and their combinations (calculated in linear scale then expressed in logarithmic scale) at each phase, (<b>a</b>) VH, (<b>b</b>) VV, (<b>c</b>) HH, (<b>d</b>) HH/VV, (<b>e</b>) VV/VH, (<b>f</b>) HH/VH, (<b>g</b>) HH × VV, (<b>h</b>) VV × VH, (<b>i</b>) HH × VH, (<b>j</b>) HH − VV, (<b>k</b>) VV − VH, (<b>l</b>) HH − VH, (<b>m</b>) HH + VV, (<b>n</b>) VV + VH, (<b>o</b>) HH + VH, (<b>p</b>) VH/(2VH + VV + HH). Red lines represent possible division values to separate at least two interquartile ranges (grey part of boxes).</p>
Full article ">Figure 8
<p>A decision tree classifier. Thresholds of VV/VH, HH/VV, and HH/VH divide SAR data into four phenological phases.</p>
Full article ">Figure 9
<p>Map of rice phenological phases spatial distribution on (<b>a</b>) 15 May, (<b>b</b>) 8 June, (<b>c</b>) 2 July, and (<b>d</b>) 26 July. Blue, green, yellow, and red colors, respectively, represent transplanting, vegetative, reproductive, and maturity phases of rice plants in the study area.</p>
Full article ">Figure 9 Cont.
<p>Map of rice phenological phases spatial distribution on (<b>a</b>) 15 May, (<b>b</b>) 8 June, (<b>c</b>) 2 July, and (<b>d</b>) 26 July. Blue, green, yellow, and red colors, respectively, represent transplanting, vegetative, reproductive, and maturity phases of rice plants in the study area.</p>
Full article ">Figure 10
<p>Evolution of observables (training dataset) provided by the eigenvalue/vector decomposition of the coherency matrix versus phenology, (<b>a</b>) Entropy, (<b>b</b>) Anisotropy, (<b>c</b>) Dominant alpha angle (<math display="inline"> <semantics> <mrow> <msub> <mi mathvariant="sans-serif">α</mi> <mn>1</mn> </msub> </mrow> </semantics> </math>, alpha of the dominant scattering mechanism). Red lines represent possible division values to separate at least two interquartile ranges (grey part of boxes).</p>
Full article ">Figure 11
<p>Phenology decision tree. Thresholds of anisotropy, entropy, and dominant alpha angle divide SAR data into four phenological phases.</p>
Full article ">Figure 12
<p>Gradual temporal change of HH/VH data on subdivisions of vegetative and reproductive phase.</p>
Full article ">
20 pages, 40917 KiB  
Article
Estimation of Global Vegetation Productivity from Global LAnd Surface Satellite Data
by Tao Yu, Rui Sun, Zhiqiang Xiao, Qiang Zhang, Gang Liu, Tianxiang Cui and Juanmin Wang
Remote Sens. 2018, 10(2), 327; https://doi.org/10.3390/rs10020327 - 22 Feb 2018
Cited by 84 | Viewed by 10046
Abstract
Accurately estimating vegetation productivity is important in research on terrestrial ecosystems, carbon cycles and climate change. Eight-day gross primary production (GPP) and annual net primary production (NPP) are contained in MODerate Resolution Imaging Spectroradiometer (MODIS) products (MOD17), which are considered the first operational [...] Read more.
Accurately estimating vegetation productivity is important in research on terrestrial ecosystems, carbon cycles and climate change. Eight-day gross primary production (GPP) and annual net primary production (NPP) are contained in MODerate Resolution Imaging Spectroradiometer (MODIS) products (MOD17), which are considered the first operational datasets for monitoring global vegetation productivity. However, the cloud-contaminated MODIS leaf area index (LAI) and Fraction of Photosynthetically Active Radiation (FPAR) retrievals may introduce some considerable errors to MODIS GPP and NPP products. In this paper, global eight-day GPP and eight-day NPP were first estimated based on Global LAnd Surface Satellite (GLASS) LAI and FPAR products. Then, GPP and NPP estimates were validated by FLUXNET GPP data and BigFoot NPP data and were compared with MODIS GPP and NPP products. Compared with MODIS GPP, a time series showed that estimated GLASS GPP in our study was more temporally continuous and spatially complete with smoother trajectories. Validated with FLUXNET GPP and BigFoot NPP, we demonstrated that estimated GLASS GPP and NPP achieved higher precision for most vegetation types. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flowchart of gross primary production (GPP) and net primary production (NPP) estimation and validation. FPAR: Fraction of Photosynthetically Active Radiation; LAI: Leaf Area Index; LUE: Light Use Efficiency; PAR: Photosynthetically Active Radiation; APAR: Absorbed Photosynthetically Active Radiation; DEM: Digital Elevation Model.</p>
Full article ">Figure 2
<p>MODerate Resolution Imaging Spectroradiometer (MODIS) International Geosphere Biosphere Program (IGBP) land-cover and location of FLUXNET sites and BigFoot sites.</p>
Full article ">Figure 3
<p>Global 1 km GPP and NPP in 2004, 2008 and 2012: (<b>a</b>) global GPP in 2004; (<b>b</b>) global NPP in 2004; (<b>c</b>) global GPP in 2008; (<b>d</b>) global NPP in 2008; (<b>e</b>) global GPP in 2012; (<b>f</b>) global NPP in 2012.</p>
Full article ">Figure 4
<p>Variations in global GPP and NPP from 2004 to 2012: (<b>a</b>) variation in GPP from 2004 to 2008; (<b>b</b>) variation in NPP from 2004 to 2008; (<b>c</b>) variation in GPP from 2008 to 2012; (<b>d</b>) variation in NPP from 2008 to 2012.</p>
Full article ">Figure 5
<p>Global GPP and NPP in 2004, 2008 and 2012 estimated using Global LAnd Surface Satellite (GLASS) data.</p>
Full article ">Figure 6
<p>Global mean and standard deviations of GPP and NPP for all vegetated land cover types: (<b>a</b>) global mean and standard deviations of GPP; (<b>b</b>) global mean and standard deviations of NPP.</p>
Full article ">Figure 7
<p>Variation in global total GPP and NPP for all vegetated land cover types: (<b>a</b>) variation in global GPP; (<b>b</b>) variation in global NPP.</p>
Full article ">Figure 8
<p>Seasonal variation in the estimated GLASS GPP, FLUXNET GPP, MOD17 C05 GPP and MOD17 C55 GPP for several sites with different vegetation types.</p>
Full article ">Figure 9
<p>Validation of estimated GLASS GPP against FLUXNET GPP.</p>
Full article ">Figure 10
<p>MODIS C05 GPP validation against FLUXNET GPP.</p>
Full article ">Figure 11
<p>MODIS C55 GPP validation against FLUXNET GPP.</p>
Full article ">Figure 12
<p>Validation of estimated GLASS NPP and MODIS NPP against BigFoot NPP: (<b>a</b>) validation of estimated GLASS NPP against BigFoot NPP; (<b>b</b>) validation of MODIS NPP against BigFoot NPP. <span class="html-italic">x</span> is the average of BigFoot NPP, the years being averaged are shown in <a href="#remotesensing-10-00327-t001" class="html-table">Table 1</a>; <span class="html-italic">y</span> is the average of NPP in 2004, 2008 and 2012.</p>
Full article ">
20 pages, 7741 KiB  
Article
Simulation and Analysis of the Topographic Effects on Snow-Free Albedo over Rugged Terrain
by Dalei Hao, Jianguang Wen, Qing Xiao, Shengbiao Wu, Xingwen Lin, Baocheng Dou, Dongqin You and Yong Tang
Remote Sens. 2018, 10(2), 278; https://doi.org/10.3390/rs10020278 - 11 Feb 2018
Cited by 34 | Viewed by 6545
Abstract
Topography complicates the modeling and retrieval of land surface albedo due to shadow effects and the redistribution of incident radiation. Neglecting topographic effects may lead to a significant bias when estimating land surface albedo over a single slope. However, for rugged terrain, a [...] Read more.
Topography complicates the modeling and retrieval of land surface albedo due to shadow effects and the redistribution of incident radiation. Neglecting topographic effects may lead to a significant bias when estimating land surface albedo over a single slope. However, for rugged terrain, a comprehensive and systematic investigation of topographic effects on land surface albedo is currently ongoing. Accurately estimating topographic effects on land surface albedo over a rugged terrain presents a challenge in remote sensing modeling and applications. In this paper, we focused on the development of a simplified estimation method for snow-free albedo over a rugged terrain at a 1-km scale based on a 30-m fine-scale digital elevation model (DEM). The proposed method was compared with the radiosity approach based on simulated and real DEMs. The results of the comparison showed that the proposed method provided adequate computational efficiency and satisfactory accuracy simultaneously. Then, the topographic effects on snow-free albedo were quantitatively investigated and interpreted by considering the mean slope, subpixel aspect distribution, solar zenith angle, and solar azimuth angle. The results showed that the more rugged the terrain and the larger the solar illumination angle, the more intense the topographic effects were on black-sky albedo (BSA). The maximum absolute deviation (MAD) and the maximum relative deviation (MRD) of the BSA over a rugged terrain reached 0.28 and 85%, respectively, when the SZA was 60° for different terrains. Topographic effects varied with the mean slope, subpixel aspect distribution, SZA and SAA, which should not be neglected when modeling albedo. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Rugged terrain with a large number of subpixel slopes; (<b>b</b>) Virtually-smoothed single slope.</p>
Full article ">Figure 2
<p>The Tibetan Plateau and the study area.</p>
Full article ">Figure 3
<p>Distributions of slope (<b>a</b>,<b>c</b>,<b>e</b>) and aspect (<b>b</b>,<b>d</b>,<b>f</b>) within a 1-km pixel under real DEMs with different mean slopes: (<b>a</b>,<b>b</b>) 10°; (<b>c</b>,<b>d</b>) 20°; and (<b>e</b>,<b>f</b>) 30°. In the legends of (<b>b</b>,<b>d</b>,<b>f</b>), N, NE, E, SE, S, SW, W and NW stand for north, northwest, east, southeast, south, southwest, west and northwest, respectively. In the north, the SAA is 0°. The SAA gradually increases in a clockwise rotation of the determined direction, until the SAA is 360° (i.e., when the SAA rotated back to its original north position).</p>
Full article ">Figure 4
<p>(<b>a</b>) Leaf reflectance, leaf transmittance and soil reflectance; (<b>b</b>) Normalized spectral irradiance curve.</p>
Full article ">Figure 5
<p>Scatterplots between the reference and the modeled BSAs over (<b>a</b>) simulated DEMs and (<b>b</b>) real DEMs.</p>
Full article ">Figure 6
<p>Hemispheric distribution of BSA under different SZAs and real 1-km DEMs: (<b>a</b>) flat terrain; (<b>b</b>) dem-10-1; (<b>c</b>) dem-20-1; and (<b>d</b>) dem-30-1. The radial coordinate is the SZA, and the angular coordinate is the SAA. The red line represents the north-south line; the backward side represents the northern aspect (i.e., where SAA is equal to 0°), and the forward side represents the southern aspect (i.e., where SAA is equal to 180°).</p>
Full article ">Figure 7
<p>BSA variation with a mean slope under different SZAs: (<b>a</b>) 0°; (<b>b</b>) 30°; and (<b>c</b>) 60°. The colors refer to the density of points (from highest (red) to lowest (blue)).</p>
Full article ">Figure 8
<p>BSA variations with subpixel aspect distributions for terrains with different mean slopes: (<b>a</b>) 10°; (<b>b</b>) 20°; and (<b>c</b>) 30°. The SAA is 150°.</p>
Full article ">Figure 9
<p>Maps of mean slope (<b>a</b>) within each 1-km pixel and the spatial distributions of BSA with different SZAs: (<b>b</b>) 0°; (<b>c</b>) 30°; and (<b>d</b>) 60°.</p>
Full article ">Figure 10
<p>BSA variation with SZA for terrains with different mean slopes: (<b>a</b>) 0°; (<b>b</b>) 10°; (<b>c</b>) 20°; and (<b>d</b>) 30°. The SAA is 150°.</p>
Full article ">Figure 11
<p>BSA variations with the SAA under different SZAs: (<b>a</b>) 0°; (<b>b</b>) 30°; (<b>c</b>) 45°; and (<b>d</b>) 60°.</p>
Full article ">Figure 12
<p>BSA variations with SAA for terrains with different mean slopes: (<b>a</b>) 10°; (<b>b</b>) 20° and (<b>c</b>) 30°. The SZA is 30°.</p>
Full article ">
28 pages, 4761 KiB  
Article
Impacts of Leaf Age on Canopy Spectral Signature Variation in Evergreen Chinese Fir Forests
by Qiaoli Wu, Conghe Song, Jinling Song, Jindi Wang, Shaoyuan Chen and Bo Yu
Remote Sens. 2018, 10(2), 262; https://doi.org/10.3390/rs10020262 - 8 Feb 2018
Cited by 13 | Viewed by 5117
Abstract
Significant gaps exist in our knowledge of the impact of leaf aging on canopy signal variability, which limits our understanding of vegetation status based on remotely sensed data. To understand the effects of leaf aging at the leaf and canopy scales, a combination [...] Read more.
Significant gaps exist in our knowledge of the impact of leaf aging on canopy signal variability, which limits our understanding of vegetation status based on remotely sensed data. To understand the effects of leaf aging at the leaf and canopy scales, a combination of field, remote-sensing and physical modeling techniques was adopted to assess the canopy spectral signals of evergreen Cunninghamia forests. We observed an approximately 10% increase in Near-Infrared (NIR) reflectance for new leaves and a 35% increase in NIR transmittance for mature leaves from May to October. When variations in leaf optical properties (LOPs) of only mature leaves, or both new and mature leaves were considered, the Geometric Optical and Radiative Transfer (GORT) model-simulated canopy reflectance trajectory was more consistent with Landsat observations (R2 increased from 0.37 to 0.82~0.89 for NIR reflectance, and from 0.35 to 0.67~0.88 for EVI2, with a small RMSE (0.01 to 0.02)). This study highlights the importance of leaf age on leaf spectral signatures, and provides evidence of age-dependent LOPs that have important impacts on canopy reflectance in the NIR band and EVI2, which are used to monitor canopy dynamics and productivity, with important implications for RS and forest ecosystem ecology. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>These four study plots are covered by <span class="html-italic">Cunninghamia lanceolata</span> (also known as Chinese fir) plantations, which were replanted after clear-cutting.</p>
Full article ">Figure 2
<p>The four scene components used in the geometric optical (GO) model. <span class="html-italic">C</span> is the sunlit tree crown; <span class="html-italic">T</span> is the shaded tree crown; <span class="html-italic">G</span> is the sunlit background; and <span class="html-italic">Z</span> is the shaded background.</p>
Full article ">Figure 3
<p>Flow chart of LOPs retrieval and application method.</p>
Full article ">Figure 4
<p>Results for the first (<b>A</b>) and second (<b>B</b>) global sensitivity analyses of the GORT model in the third (red) and fourth (NIR) Landsat bands for the following parameters: LAI (leaf area index), <math display="inline"> <semantics> <mrow> <mi>h</mi> <mn>1</mn> </mrow> </semantics> </math> (lower boundary of canopy center height), <math display="inline"> <semantics> <mrow> <mi>h</mi> <mn>2</mn> </mrow> </semantics> </math> (upper boundary of canopy center height), <math display="inline"> <semantics> <mi>λ</mi> </semantics> </math> (tree stem density (trees/ha)), <span class="html-italic">r</span> (crown radius), <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>L</mi> </msub> </mrow> </semantics> </math> (leaf reflectance), <math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mi>L</mi> </msub> </mrow> </semantics> </math> (leaf transmittance), and <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>G</mi> </msub> </mrow> </semantics> </math> (background reflectance).</p>
Full article ">Figure 5
<p>Single order sensitivity analysis of the impacts of sensitive parameters on GORT model outputs for the red band (<b>A</b>) and the NIR band (<b>B</b>). Only one of the model sensitive parameters was adjusted each time to study variations in the model output spectral signatures.</p>
Full article ">Figure 6
<p>Leaf reflectance (<math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>L</mi> </msub> </mrow> </semantics> </math>, <b>A1</b>–<b>C1</b>) and transmittance (<math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mi>L</mi> </msub> </mrow> </semantics> </math>, <b>A2</b>–<b>C2</b>) for Chinese fir leaves at different ages (0–3 a) at three (green, red and NIR) short-wave bands. Leaves were collected on 5 May 2017.</p>
Full article ">Figure 7
<p>Field spectral curve data (mean ± 1 s.d.) in shortwave bands for new (<b>A1</b>–<b>E1</b>) and mature (<b>A2</b>–<b>E2</b>) needle samples (<span class="html-italic">n</span>) from Chinese fir collected on 5 May, 24 June, 28 July, 14 September and 13 October 2017. From May to October, <span class="html-italic">n</span> = 22, 15, 26, 30 and 39 respectively, and each sample includes 5 to 8 leaves. Reflectance data (<math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>L</mi> </msub> </mrow> </semantics> </math>) for each month are presented as the lower set of curves within each plot, while transmittance data (<math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mi>L</mi> </msub> </mrow> </semantics> </math>) are presented as the upper set of curves. Absorption characteristics are depicted based on the area between the upper and lower set of curves. SD values for new leaf <math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mi>L</mi> </msub> </mrow> </semantics> </math> are depicted in the upper and lower dash lines, to avoid overlap with <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>L</mi> </msub> </mrow> </semantics> </math>, while SD values for all other <math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>L</mi> </msub> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mi>L</mi> </msub> </mrow> </semantics> </math> measurements are depicted in gray buffed areas.</p>
Full article ">Figure 8
<p>Retrieved results for leaf reflectance (<math display="inline"> <semantics> <mrow> <msub> <mi>r</mi> <mi>L</mi> </msub> </mrow> </semantics> </math>) and transmittance (<math display="inline"> <semantics> <mrow> <msub> <mi>t</mi> <mi>L</mi> </msub> </mrow> </semantics> </math>) at the canopy scale in the NIR band (<b>A1</b>,<b>A2</b>) and RED band (<b>B1</b>,<b>B2</b>).</p>
Full article ">Figure 9
<p>Seasonal variations in the leaf properties of the canopy, including (<b>A</b>) total leaf area: average LAI from stand age 22 to 33 (2005 to 2015); (<b>B</b>) leaf proportion: the percentage of new leaves and mature leaves in the canopy measured by Zhongkun et al. [<a href="#B50-remotesensing-10-00262" class="html-bibr">50</a>]; (<b>C</b>) weights for new leaves (<span class="html-italic">w</span>1) and mature leaves (<span class="html-italic">w</span>2) from stand age 22 to 33 (2005 to 2015).</p>
Full article ">Figure 10
<p>Validation of estimated new leaf LOPs in the NIR band (<b>A1</b>) and red band (<b>A2</b>) and the constructed mature leaf LOPs trajectories (<b>B1</b>,<b>B2</b>).</p>
Full article ">Figure 11
<p>Aging effects of new leaves and mature leaves on seasonality of canopy signal trajectory in NIR band (<b>A1</b>), red band (<b>B1</b>) and EVI2 (<b>C1</b>). With the exception of the differences in leaf optical parameters, all other input parameters for the GORT model are the same in three circumstances. We evaluated the simulated canopy signature with Landsat observations (<b>A2</b>–<b>C2</b>). Both the simulated results and Landsat observations are at a monthly step with the mean value and s.d. (from 2005 to 2015).</p>
Full article ">Figure A1
<p>Measured parameters for sample trees selected to build regression relationships for the structural parameters of the tree crown and other measures, including tree height and DBH.</p>
Full article ">Figure A2
<p>(<b>A</b>) Measured crown width (2<span class="html-italic">R</span>) for 40 sample trees with DBH in the north-south direction (CW<sub>NS</sub>) and east-west direction (CW<sub>EW</sub>); (<b>B</b>) Regression relationships between DBH and crown width (2<span class="html-italic">R</span>).</p>
Full article ">Figure A3
<p>(<b>A</b>) Measured height values for 40 sample trees with increasing DBH. The measured data include tree height (<span class="html-italic">H</span>1), height of crown center (<span class="html-italic">h</span>) and height under crown (<span class="html-italic">H</span>2); Regression relationships between DBH and tree height (<b>B</b>) and between tree height and crown center height (<b>C</b>).</p>
Full article ">Figure A4
<p>The vacuum side of one group of leaf leaves prepared for spectral measurement. Leaf samples were collected on 15 July 2016.</p>
Full article ">Figure A5
<p>Flowchart of LAI field data processing procedures. Before January 2007, monthly LAI was measured using a CI-110 instrument, and a DHP instrument was used afterwards.</p>
Full article ">
17 pages, 20645 KiB  
Article
Estimation of High Spatial-Resolution Clear-Sky Land Surface-Upwelling Longwave Radiation from VIIRS/S-NPP Data
by Shugui Zhou and Jie Cheng
Remote Sens. 2018, 10(2), 253; https://doi.org/10.3390/rs10020253 - 7 Feb 2018
Cited by 12 | Viewed by 4035
Abstract
Surface-upwelling longwave radiation (LWUP) is an important component of the surface radiation budget. Under the general framework of the hybrid method, the linear models and the multivariate adaptive regression spline (MARS) models are developed to estimate the 750 m instantaneous clear-sky LWUP from [...] Read more.
Surface-upwelling longwave radiation (LWUP) is an important component of the surface radiation budget. Under the general framework of the hybrid method, the linear models and the multivariate adaptive regression spline (MARS) models are developed to estimate the 750 m instantaneous clear-sky LWUP from the top-of-atmosphere (TOA) radiance of the Visible Infrared Imaging Radiometer Suite (VIIRS) channels M14, M15, and M16. Comprehensive radiative transfer simulations are conducted to generate a huge amount of representative samples, from which the linear model and the MARS model are derived. The two models developed are validated by the field measurements collected from seven sites in the Surface Radiation Budget Network (SURFRAD). The bias and root-mean-square error (RMSE) of the linear models are −4.59 W/m2 and 16.15 W/m2, whereas those of the MARS models are −5.23 W/m2 and 16.38 W/m2, respectively. The linear models are preferable for the production of the operational LWUP product due to its higher computational efficiency and acceptable accuracy. The LWUP estimated by the linear models developed from VIIRS is compared to that retrieved from the Moderate Resolution Imaging Spectroradiometer (MODIS). They agree well with each other with bias and RMSE of −0.15 W/m2 and 25.24 W/m2 respectively. This is the first time that the hybrid method has been applied to globally estimate clear-sky LWUP from VIIRS data. The good performance of the developed hybrid method and consistency between VIIRS LWUP and MODIS LWUP indicate that the hybrid method is promising for producing the long-term high spatial resolution environmental data record (EDR) of LWUP. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Relative spectral responses of Visible Infrared Imaging Radiometer Suite (VIIRS) channels M14, M15 and M16, and Moderate Resolution Imaging Spectroradiometer (MODIS) channels 29, 31 and 32. The gray line represents the transmittance of the 1976 U.S. Standard Atmosphere.</p>
Full article ">Figure 2
<p>Flowchart for developing the hybrid method.</p>
Full article ">Figure 3
<p>Validation results of the linear model at SURFRAD sites. (<b>a</b>) Bondville_IL; (<b>b</b>) Boulder_CO; (<b>c</b>) Desert_Rock_NV; (<b>d</b>) Fort_Peck_MT; (<b>e</b>) Goodwin_Creek_MS; (<b>f</b>) Penn_State_PA; (<b>g</b>) Sioux_Falls_SD.</p>
Full article ">Figure 4
<p>Validation results of the MARS model at SURFRAD sites. (<b>a</b>) Bondville_IL. (<b>b</b>) Boulder_CO. (<b>c</b>) Desert_Rock_NV. (<b>d</b>) Fort_Peck_MT. (<b>e</b>) Goodwin_Creek_MS. (<b>f</b>) Penn_State_PA. (<b>g</b>) Sioux_Falls_SD.</p>
Full article ">Figure 4 Cont.
<p>Validation results of the MARS model at SURFRAD sites. (<b>a</b>) Bondville_IL. (<b>b</b>) Boulder_CO. (<b>c</b>) Desert_Rock_NV. (<b>d</b>) Fort_Peck_MT. (<b>e</b>) Goodwin_Creek_MS. (<b>f</b>) Penn_State_PA. (<b>g</b>) Sioux_Falls_SD.</p>
Full article ">Figure 5
<p>Distribution of LWUPs derived from VIIRS and MODIS images.</p>
Full article ">Figure 6
<p>The scatterplot of retrieved LWUP from VIIRS and that retrieved from MODIS.</p>
Full article ">Figure 7
<p>Validation results of the linear model at the Atmospheric Radiation Measurement (ARM) program Southern Great Plains (SGP) C1 site. (<b>a</b>) Pixels that were identified as clear sky by the VIIRS cloud mask whereas cloudy by ground-based Lidar; (<b>b</b>) pixels that were identified as clear sky by both the VIIRS Cloud Mask and the Lidar.</p>
Full article ">Figure 8
<p>The relationships between Broadband Emissivity (BBE) versus bias of retrieved LWUP.</p>
Full article ">
26 pages, 3837 KiB  
Article
Evaluating the Performance of the SCOPE Model in Simulating Canopy Solar-Induced Chlorophyll Fluorescence
by Jiaochan Hu, Xinjie Liu, Liangyun Liu and Linlin Guan
Remote Sens. 2018, 10(2), 250; https://doi.org/10.3390/rs10020250 - 6 Feb 2018
Cited by 27 | Viewed by 6386
Abstract
The SCOPE (soil canopy observation of photochemistry and energy fluxes) model has been widely used to interpret solar-induced chlorophyll fluorescence (SIF) and investigate the SIF-photosynthesis links at different temporal and spatial scales in recent years. In the SCOPE model, the fluorescence quantum efficiency [...] Read more.
The SCOPE (soil canopy observation of photochemistry and energy fluxes) model has been widely used to interpret solar-induced chlorophyll fluorescence (SIF) and investigate the SIF-photosynthesis links at different temporal and spatial scales in recent years. In the SCOPE model, the fluorescence quantum efficiency in dark-adapted conditions (FQE) for Photosystem II (fqe2) and Photosystem I (fqe1) were two key parameters of SIF emission, which have always been parameterized as fixed values derived from laboratory measurements. To date, only a few studies have focused on evaluating the SCOPE model for SIF interpretation, and the variation of FQE values in the field remains controversial. In this study, the accuracy of the SCOPE model to simulate the canopy SIF was investigated using diurnal experiments on winter wheat. First, ten diurnal experiments were conducted on winter wheat, and the canopy SIF emissions and the SCOPE model’s input parameters were directly measured or indirectly retrieved from the spectral radiances, gross primary productivity (GPP) data, and meteorological records. Second, the SCOPE-simulated SIF emissions with fixed FQE values were evaluated using the observed canopy SIF data. The results show that the SCOPE model can reliably interpret the diurnal cycles of SIF variation and provide acceptable results of SIF simulations at the O2-B (SIFB) and O2-A (SIFA) bands with RRMSEs of 24.35% and 23.67%, respectively. However, the SCOPE-simulated SIFB and SIFA still contained large systematical deviations at some growth stages of wheat, and the seasonal cycles of the ratio between SIFB and SIFA (SIFA/SIFB) cannot be credibly reproduced. Finally, the SCOPE-simulated SIF emissions with variable FQE values were evaluated using the observed canopy SIF data. The simulating accuracy of SIFB and SIFA can be improved greatly using variable FQE values, and the SCOPE simulations track well with the seasonal SIFA/SIFB values with an RRMSE of 20.63%. The results indicated a clear seasonal pattern of FQE values for unbiased SIF simulation: from the erecting to the flowering stage of wheat, the ratio of fqe1 to fqe2 (fqe1/fqe2) gradually increased from 0.05–0.1 to 0.3–0.5, while the fqe2 value decreased from 0.013 to 0.007. Our quantitative results of the model assessment and the FQE adjustment support the use of the SCOPE model as a powerful tool for interpreting the SIF emissions and can serve as a significant reference for future applications of the SCOPE model. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Diurnal observations of meteorological parameters between 7:00 to 19:00 made during ten experiments in 2015 and 2016: (<b>a</b>) half-hour air temperature and VPD observations and (<b>b</b>) half-hour PAR observations.</p>
Full article ">Figure 2
<p>Schematic overview of the LIDFa inversion procedure.</p>
Full article ">Figure 3
<p>Schematic overview of the V<sub>cmo</sub> inversion procedure.</p>
Full article ">Figure 4
<p>Schematic overview of SIF simulations and model evaluation.</p>
Full article ">Figure 5
<p>The simulated and measured reflectance spectra and their residuals at 10:00 for every fieldwork day in 2015 and 2016.</p>
Full article ">Figure 6
<p>Root mean squared error (RMSE) values between the simulated and measured reflectance spectra for 106 spectral measurements in 2015 and 2016.</p>
Full article ">Figure 7
<p>The diurnal cycles of simulated and measured GPP at half-hour intervals between 7:00 and 19:00 on ten fieldwork days in 2015 and 2016.</p>
Full article ">Figure 8
<p>The correlation and relative root mean square error (RRMSE) value between simulated and measured GPP for all half-hour flux observations during ten experiments in 2015 and 2016.</p>
Full article ">Figure 9
<p>The diurnal cycles of simulated SIF with fixed FQE, compared to the measured SIF at O<sub>2</sub>-B and O<sub>2</sub>-A bands on ten fieldwork days in 2015 and 2016.</p>
Full article ">Figure 10
<p>The correlation and relative root mean square error (RRMSE) values between simulated and measured SIF<sub>A</sub>/SIF<sub>B</sub> with fixed FQE: (<b>a</b>) the correlation and RRMSE for 106 spectral measurements and (<b>b</b>) the RRMSE values for each fieldwork day in 2015 and 2016.</p>
Full article ">Figure 11
<p>The seasonal cycles of fqe2 and fqe1/fqe2 on ten fieldwork days in 2015 and 2016.</p>
Full article ">Figure 12
<p>The correlation and relative root mean square error (RRMSE) values between simulated and measured SIF<sub>A</sub> and SIF<sub>B</sub> with variable FQE: (<b>a</b>) the correlation and RRMSE for 106 spectral measurements and (<b>b</b>) the RRMSE values for each fieldwork day in 2015 and 2016.</p>
Full article ">Figure 13
<p>The correlation and relative root mean square error (RRMSE) values between simulated and measured SIF<sub>A</sub>/SIF<sub>B</sub> with variable FQE: (<b>a</b>) the correlation and RRMSE for 106 spectral measurements and (<b>b</b>) the RRMSE values for each fieldwork day in 2015 and 2016.</p>
Full article ">
15 pages, 5099 KiB  
Article
Retrieval of Aerosol Optical Depth in the Arid or Semiarid Region of Northern Xinjiang, China
by Xinpeng Tian, Sihai Liu, Lin Sun and Qiang Liu
Remote Sens. 2018, 10(2), 197; https://doi.org/10.3390/rs10020197 - 29 Jan 2018
Cited by 24 | Viewed by 4831
Abstract
Satellite remote sensing has been widely used to retrieve aerosol optical depth (AOD), which is an indicator of air quality as well as radiative forcing. The dark target (DT) algorithm is applied to low reflectance areas, such as dense vegetation, and the deep [...] Read more.
Satellite remote sensing has been widely used to retrieve aerosol optical depth (AOD), which is an indicator of air quality as well as radiative forcing. The dark target (DT) algorithm is applied to low reflectance areas, such as dense vegetation, and the deep blue (DB) algorithm is adopted for bright-reflecting regions. However, both DT and DB algorithms ignore the effect of surface bidirectional reflectance. This paper provides a method for AOD retrieval in arid or semiarid areas, in which the key points are the accurate estimation of surface reflectance and reasonable assumptions of the aerosol model. To reduce the uncertainty in surface reflectance, a minimum land surface reflectance database at the spatial resolution of 500 m for each month was constructed based on the moderate-resolution imaging spectroradiometer (MODIS) surface reflectance product. Furthermore, a bidirectional reflectance distribution function (BRDF) correction model was adopted to compensate for the effect of surface reflectance anisotropy. The aerosol parameters, including AOD, single scattering albedo, asymmetric factor, Ångström exponent and complex refractive index, are determined based on the observation of two sunphotometers installed in northern Xinjiang from July to August 2014. The AOD retrieved from the MODIS images was validated with ground-based measurements and the Terra-MODIS aerosol product (MOD04). The 500 m AOD retrieved from the MODIS showed high consistency with ground-based AOD measurements, with an average correlation coefficient of ~0.928, root mean square error (RMSE) of ~0.042, mean absolute error (MAE) of ~0.032, and the percentage falling within the expected error (EE) of the collocations is higher than that for the MOD04 DB product. The results demonstrate that the new AOD algorithm is more suitable to represent aerosol conditions over Xinjiang than the DB standard product. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Map of the study area and the two ground-observed sites.</p>
Full article ">Figure 2
<p>Flowchart of the aerosol retrieval algorithm in the study.</p>
Full article ">Figure 3
<p>Example of a pre-calculated minimum land surface reflectance (MLSR) database using 5 years of MOD09A1 at 500 m resolution for July. (<b>a</b>): the surface reflectance at blue band; (<b>b</b>): the solar zenith angle; (<b>c</b>): the viewing zenith angle; (<b>d</b>): the relative azimuth angle.</p>
Full article ">Figure 4
<p>Time series of aerosol parameters of single scattering albedo (SSA) and g retrieved from sky radiance almucantar measurements and direct sun measurements. 15 July, 3 August, 4 August, and 21 August are the start date, the stable weather data, the maximum AOD date, and the end date, respectively.</p>
Full article ">Figure 5
<p>MODIS false-color images for the northern Xinjiang area (R, G, B = 2, 1, 4). (<b>a</b>): 11 July 2014; (<b>b</b>): 15 July 2014; (<b>c</b>): 12 August 2014; (<b>d</b>): 21 August 2014.</p>
Full article ">Figure 5 Cont.
<p>MODIS false-color images for the northern Xinjiang area (R, G, B = 2, 1, 4). (<b>a</b>): 11 July 2014; (<b>b</b>): 15 July 2014; (<b>c</b>): 12 August 2014; (<b>d</b>): 21 August 2014.</p>
Full article ">Figure 6
<p>Retrieved AOD for the northern Xinjiang area. (<b>a</b>): 11 July 2014; (<b>b</b>): 15 July 2014; (<b>c</b>): 12 August 2014; (<b>d</b>): 21 August 2014.</p>
Full article ">Figure 7
<p>Distribution of MODIS 10-km deep blue (DB) AOD products for the northern Xinjiang area. (<b>a</b>): 11 July 2014; (<b>b</b>): 15 July 2014; (<b>c</b>): 12 August 2014; (<b>d</b>): 21 August 2014.</p>
Full article ">Figure 8
<p>Validation of retrieved AOD from the (<b>a</b>) new algorithm (500 m), (<b>b</b>) MOD04 C6 DB algorithm (10 km), and (<b>c</b>) the new algorithm without angle normalization (500 m) against the ground-based sunphotometer AOD measurements. The black dashed lines are the EE lines, the black solid lines are the 1:1 line and red solid lines are the regression lines.</p>
Full article ">Figure 9
<p>The variation of surface reflectance during the same month in different years near the ground sites. (<b>a</b>) Dahuangshan; (<b>b</b>) Wucaiwan. The error bars represent the maximum absolute error compared with the average value for 5 years.</p>
Full article ">
22 pages, 20862 KiB  
Article
Estimating Surface Downward Shortwave Radiation over China Based on the Gradient Boosting Decision Tree Method
by Lu Yang, Xiaotong Zhang, Shunlin Liang, Yunjun Yao, Kun Jia and Aolin Jia
Remote Sens. 2018, 10(2), 185; https://doi.org/10.3390/rs10020185 - 26 Jan 2018
Cited by 61 | Viewed by 7167
Abstract
Downward shortwave radiation (DSR) is an essential parameter in the terrestrial radiation budget and a necessary input for models of land-surface processes. Although several radiation products using satellite observations have been released, coarse spatial resolution and low accuracy limited their application. It is [...] Read more.
Downward shortwave radiation (DSR) is an essential parameter in the terrestrial radiation budget and a necessary input for models of land-surface processes. Although several radiation products using satellite observations have been released, coarse spatial resolution and low accuracy limited their application. It is important to develop robust and accurate retrieval methods with higher spatial resolution. Machine learning methods may be powerful candidates for estimating the DSR from remotely sensed data because of their ability to perform adaptive, nonlinear data fitting. In this study, the gradient boosting regression tree (GBRT) was employed to retrieve DSR measurements with the ground observation data in China collected from the China Meteorological Administration (CMA) Meteorological Information Center and the satellite observations from the Advanced Very High Resolution Radiometer (AVHRR) at a spatial resolution of 5 km. The validation results of the DSR estimates based on the GBRT method in China at a daily time scale for clear sky conditions show an R2 value of 0.82 and a root mean square error (RMSE) value of 27.71 W·m−2 (38.38%). These values are 0.64 and 42.97 W·m−2 (34.57%), respectively, for cloudy sky conditions. The monthly DSR estimates were also evaluated using ground measurements. The monthly DSR estimates have an overall R2 value of 0.92 and an RMSE of 15.40 W·m−2 (12.93%). Comparison of the DSR estimates with the reanalyzed and retrieved DSR measurements from satellite observations showed that the estimated DSR is reasonably accurate but has a higher spatial resolution. Moreover, the proposed GBRT method has good scalability and is easy to apply to other parameter inversion problems by changing the parameters and training data. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Spatial distribution of the radiation sites provided by the China Meteorological Administration (CMA) Meteorological Information Center.</p>
Full article ">Figure 2
<p>The main procedures of the gradient boosting regression tree (GBRT) method.</p>
Full article ">Figure 3
<p>Flowchart of the GBRT method.</p>
Full article ">Figure 4
<p>Artificial neural network (ANN) structure used in this study.</p>
Full article ">Figure 5
<p>(<b>a</b>) Evaluation results of the training set’s daily estimated DSR based on the GBRT-based clear sky model against ground measurements in 2001 and 2002. (<b>b</b>) Evaluation results of the validation set’s daily estimated DSR based on the GBRT-based clear sky model against ground measurements in 2003. The number in the parentheses is the percent bias or root mean square error (RMSE) value.</p>
Full article ">Figure 6
<p>(<b>a</b>) Evaluation results of the training set’s daily estimated DSR based on the GBRT-based cloudy sky model against ground measurements in 2001 and 2002. (<b>b</b>) Evaluation results of the validation set’s daily estimated DSR based on the GBRT-based cloudy sky model against ground measurements in 2003. The number in the parentheses is the percent bias or RMSE value.</p>
Full article ">Figure 7
<p>Validation results of the estimated daily DSR based on the GBRT model under clear sky conditions without considering Advanced Very High Resolution Radiometer (AVHRR) channels 4 and 5 as the input variables. The number in the parentheses is the percent bias or RMSE value.</p>
Full article ">Figure 8
<p>Validation results of the estimated daily DSR based on the GBRT model without considering AVHRR channels 4 and 5 as the input variables under cloudy sky conditions. The number in the parentheses is the percent bias or RMSE value.</p>
Full article ">Figure 9
<p>(<b>a</b>) Evaluation results of the training set’s estimated monthly mean DSR based on the GBRT-based DSR model against ground measurements in 2001 and 2002. (<b>b</b>) Evaluation results of the validation set’s estimated monthly mean DSR based on the GBRT-based DSR model against ground measurements in 2003. The number in the parentheses is the percent bias or RMSE value.</p>
Full article ">Figure 10
<p>(<b>a</b>) Evaluation results of the training dataset’s daily estimated DSR based on the ANN-based clear sky model against ground measurements in 2001 and 2002. (<b>b</b>) Evaluation results of the validation dataset’s daily estimated DSR based on the ANN-based clear sky model against ground measurements in 2003. The number in the parentheses is the RMSE value.</p>
Full article ">Figure 11
<p>(<b>a</b>) Evaluation results of the training dataset’s daily estimated DSR based on the ANN-based cloudy sky model against ground measurements in 2001 and 2002. (<b>b</b>) Evaluation results of the validation dataset’s daily estimated DSR based on the ANN-based cloudy sky model against ground measurements in 2003. The number in the parentheses is the percent bias or RMSE value.</p>
Full article ">Figure 12
<p>(<b>a</b>) Evaluation results of the training set’s estimated monthly mean DSR based on the ANN-based DSR model against ground measurements in 2001 and 2002. (<b>b</b>) Evaluation results of the validation set’s estimated monthly mean DSR based on the ANN-based DSR model against ground measurements in 2003. The number in the parentheses is the percent bias or RMSE value.</p>
Full article ">Figure 13
<p>The spatial distribution of the DSR estimates from (<b>a</b>) the GBRT model, (<b>b</b>) the GEWEX-SRB, and (<b>c</b>) the MERRA in March 2003. (<b>d</b>) The differences between monthly mean DSR estimates of the GEWEX-SRB and the GBRT model (i.e., the GEWEX-SRB estimates minus the GBRT-based estimates) in March 2003. (<b>e</b>) The differences between monthly mean DSR estimates of the MERRA and the GBRT model (i.e., the MERRA estimates minus the GBRT-based estimates) in March 2003.</p>
Full article ">Figure 14
<p>Scatter plots comparing the results from (<b>a</b>) the GBRT-based DSR model, as well as the DSR products (<b>b</b>) the Global Energy and Water Cycle Experiment-Surface Radiation Budget (GEWEX-SRB) and (<b>c</b>) Modern-Era Retrospective analysis for Research and Applications <tt>(</tt>MERRA) against ground measurements in 2003. The number in the parentheses is the percent bias or RMSE value.</p>
Full article ">
15 pages, 2773 KiB  
Article
SPI-Based Analyses of Drought Changes over the Past 60 Years in China’s Major Crop-Growing Areas
by Lang Xia, Fen Zhao, Kebiao Mao, Zijin Yuan, Zhiyuan Zuo and Tongren Xu
Remote Sens. 2018, 10(2), 171; https://doi.org/10.3390/rs10020171 - 25 Jan 2018
Cited by 31 | Viewed by 5766
Abstract
This study analyzes the changes in drought patterns in China’s major crop-growing areas over the past 60 years. The analysis was done using both weather station data and Tropical Rainfall Measuring Mission (TRMM) Microwave Imager (TMI) rainfall data to calculate the Standardized Precipitation [...] Read more.
This study analyzes the changes in drought patterns in China’s major crop-growing areas over the past 60 years. The analysis was done using both weather station data and Tropical Rainfall Measuring Mission (TRMM) Microwave Imager (TMI) rainfall data to calculate the Standardized Precipitation Index (SPI). The results showed that the occurrences of extreme drought were the most serious in recent years in the Southwest China and Sichuan crop-growing areas. The Yangtze River (MLRY) and South China crop-growing areas experienced extreme droughts during 1960–1980, whereas the Northeast China and Huang–Huai–Hai crop-growing areas experienced extreme droughts around 2003. The analysis showed that the SPIs calculated by TRMM data at time scales of one, three, and six months were reliable for monitoring drought in the study regions, but for 12 months, the SPIs calculated by gauge and TRMM data showed less consistency. The analysis of the spatial distribution of droughts over the past 15 years using TMI rainfall data revealed that more than 60% of the area experienced extreme drought in 2011 over the MLRY region and in 1998 over the Huang–Huai–Hai region. The frequency of different intensity droughts presented significant spatial heterogeneity in each crop-growing region. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study areas and meteorological station locations.</p>
Full article ">Figure 2
<p>The SPI values for 1951–2013 for (<b>a</b>) Northeast China and (<b>b</b>) Huang–Huai–Hai.</p>
Full article ">Figure 3
<p>The SPI values for 1951–2013 for (<b>a</b>) MLRY and (<b>b</b>) South areas.</p>
Full article ">Figure 4
<p>The SPI values for 1951–2013 for (<b>a</b>) Sichuan basin and (<b>b</b>) Southwest China.</p>
Full article ">Figure 5
<p>Frequencies of moderate droughts, severe droughts and extreme droughts for different regions from 1998 to 2013.</p>
Full article ">Figure 6
<p>TRMM 3B43 for drought monitoring (SPI at time scale of six months) for different regions: (<b>a</b>) Northeast; (<b>b</b>) Huang–Huai–Hai; (<b>c</b>) MLRY; (<b>d</b>) South; (<b>e</b>) Sichuan basin; and (<b>f</b>) Southwest.</p>
Full article ">Figure 6 Cont.
<p>TRMM 3B43 for drought monitoring (SPI at time scale of six months) for different regions: (<b>a</b>) Northeast; (<b>b</b>) Huang–Huai–Hai; (<b>c</b>) MLRY; (<b>d</b>) South; (<b>e</b>) Sichuan basin; and (<b>f</b>) Southwest.</p>
Full article ">Figure 7
<p>Coefficient of determination R<sup>2</sup> between two datasets for difference regions: (<b>a</b>) Northeast; (<b>b</b>) Huang–Huai–Hai; (<b>c</b>) MLRY; (<b>d</b>) South; (<b>e</b>) Sichuan basin; (<b>f</b>) Southwest.</p>
Full article ">Figure 8
<p>Gauge data for drought monitoring (SPI at time scale of 6 month) for difference regions: (<b>a</b>) Northeast, (<b>b</b>) Huang–Huai–Hai, (<b>c</b>) MLRY, (<b>d</b>) South, (<b>e</b>) Sichuan basin and (<b>f</b>) Southwest.</p>
Full article ">Figure 8 Cont.
<p>Gauge data for drought monitoring (SPI at time scale of 6 month) for difference regions: (<b>a</b>) Northeast, (<b>b</b>) Huang–Huai–Hai, (<b>c</b>) MLRY, (<b>d</b>) South, (<b>e</b>) Sichuan basin and (<b>f</b>) Southwest.</p>
Full article ">
18 pages, 6996 KiB  
Article
Inter-Comparison and Evaluation of Remote Sensing Precipitation Products over China from 2005 to 2013
by Qiaolin Zeng, Yongqian Wang, Liangfu Chen, Zifeng Wang, Hao Zhu and Bin Li
Remote Sens. 2018, 10(2), 168; https://doi.org/10.3390/rs10020168 - 25 Jan 2018
Cited by 45 | Viewed by 5915
Abstract
Precipitation is a key aspect of the climate system. In this paper, the dependability of five satellite precipitation products (TRMM [Tropical Rainfall Measuring Mission] 3BV42, PERSIANN [Precipitation Estimation from Remotely Sensed Information using Artificial Neural Networks] CDR, GSMaP [Global Satellite Mapping of Precipitation] [...] Read more.
Precipitation is a key aspect of the climate system. In this paper, the dependability of five satellite precipitation products (TRMM [Tropical Rainfall Measuring Mission] 3BV42, PERSIANN [Precipitation Estimation from Remotely Sensed Information using Artificial Neural Networks] CDR, GSMaP [Global Satellite Mapping of Precipitation] RENALYSIS, CMORPH [Climate Prediction Center’s morphing technique] BLD and CMORPH_RAW) were compared with in situ measurements over China for the period of 2005 to 2013. To completely evaluate these precipitation products, the annual, seasonal and monthly precipitation averages were calculated. Overall, the Huaihe River and Qinlin mountains are shown to have heavy precipitation to the southeast and lighter precipitation to the northwest. The comparison results indicate that Gauge correction (CMORPH_BLD) improves the quality of the original satellite products (CMORPH_RAW), resulting in the higher correlation coefficient (CC), the low relative bias (BIAS) and root mean square error (RMSE). Over China, the GSMaP_RENALYSIS outperforms other products and shows the highest CC (0.91) and lowest RMSE (0.85 mm/day) and all products except for PERSIANN_CDR exhibit underestimation. GSMaP_RENALYSIS gives the highest of probability of detection (81%), critical success index (63%) and lowest false alarm ratio (36%) while TRMM3BV42 gives the highest of frequency bias index (1.00). Over Tibetan Plateau, CMORPH_RAW demonstrates the poorest performance with the biggest BIAS (4.2 mm/month) and lowest CC (0.22) in December 2013. GSMaP_RENALYSIS displays quite consistent with in situ measurements in summer. However, GSMaP_RENALYSIS and CMORPH_RAW underestimate precipitation over South China. CMORPH_BLD and TRMM3BV42 show consistent with high CC (>0.8) but relatively large RMSE in summer. Full article
Show Figures

Figure 1

Figure 1
<p>Site distribution and topography.</p>
Full article ">Figure 2
<p>Spatial features of the nine-year mean daily rainfall obtained from CMOPRH_BLD, CMORPH_RAW, TRMM3BV42, PERSIANN_CDR, GSMaP_RENALYSIS and GAUGE_STATION.</p>
Full article ">Figure 3
<p>Density-colored scatterplots of the different products against the in-situ measurements for the nine-year mean daily precipitation. The red line is the fit and the black is 1:1. (<b>a</b>) denotes CMOPRH_RAW &amp;gauge, (<b>b</b>) denotes CMOPRH_BLD &amp;gauge, (<b>c</b>) denotes TRMM3BV42 &amp;gauge, (<b>d</b>) denotes PERSIANN_CDR &amp;gauge, (<b>e</b>) denotes GSMaP_RENALYSIS &amp;gauge.</p>
Full article ">Figure 4
<p>Seasonal nine-year daily mean rainfall features of CMOPRH_BLD, CMORPH_RAW, TRMM3BV42, PERSIANN_CDR, GSMaP_RENALYSIS and GAUGE_STATION.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>c</b>) denote the CC, BIAS and RMSE values of the satellite products and rain gauge stations for the seasonal daily mean precipitation values, respectively.</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>c</b>) denote the time series of the CC, RMSE and BIAS from a variety products spanning 2005 to 2013 over China, respectively.</p>
Full article ">Figure 7
<p>Probability density function of daily rainfall for the no-rain case.</p>
Full article ">Figure 8
<p>Probability density function of the daily rainfall events with different intensities.</p>
Full article ">Figure 9
<p>The precipitation distributions of the rain gauge stations (<b>a</b>) and multiple satellite (<b>b</b>) in 2013 over southern China.</p>
Full article ">Figure 10
<p>(<b>a</b>–<b>c</b>) denote the CC, RMSE and BIAS among a variety of products for each month in 2013 over the Tibetan Plateau, respectively.</p>
Full article ">Figure 11
<p>The precipitation distributions from the rain gauge stations (<b>a</b>) and multiple satellites (<b>b</b>) in 2013 over southern China.</p>
Full article ">Figure 12
<p>(<b>a</b>–<b>c</b>) denote the CC, RMSE and BIAS values of the studied products for each month in 2013 over southern China, respectively.</p>
Full article ">
23 pages, 3567 KiB  
Article
A Multi-Scale Validation Strategy for Albedo Products over Rugged Terrain and Preliminary Application in Heihe River Basin, China
by Xingwen Lin, Jianguang Wen, Qinhuo Liu, Qing Xiao, Dongqin You, Shengbiao Wu, Dalei Hao and Xiaodan Wu
Remote Sens. 2018, 10(2), 156; https://doi.org/10.3390/rs10020156 - 24 Jan 2018
Cited by 17 | Viewed by 4893
Abstract
The issue for the validation of land surface remote sensing albedo products over rugged terrain is the scale effects between the reference albedo measurements and coarse scale albedo products, which is caused by the complex topography. This paper illustrates a multi-scale validation strategy [...] Read more.
The issue for the validation of land surface remote sensing albedo products over rugged terrain is the scale effects between the reference albedo measurements and coarse scale albedo products, which is caused by the complex topography. This paper illustrates a multi-scale validation strategy specified for coarse scale albedo validation over rugged terrain. A Mountain-Radiation-Transfer-based (MRT-based) albedo upscaling model was proposed in the process of multi-scale validation strategy for aggregating fine scale albedo to coarse scale. The simulated data of both the reference coarse scale albedo and fine scale albedo were used to assess the performance and uncertainties of the MRT-based albedo upscaling model. The results showed that the MRT-based model could reflect the albedo scale effects over rugged terrain and provided a robust solution for albedo upscaling from fine scale to coarse scale with different mean slopes and different solar zenith angles. The upscaled coarse scale albedos had the great agreements with the simulated coarse scale albedo with a Root-Mean-Square-Error (RMSE) of 0.0029 and 0.0017 for black sky albedo (BSA) and white sky albedo (WSA), respectively. Then the MRT-based model was preliminarily applied for the assessment of daily MODerate Resolution Imaging Spectroradiometer (MODIS) Albedo Collection V006 products (MCD43A3 C6) over rugged terrain. Results showed that the MRT-based model was effective and suitable for conducting the validation of MODIS albedo products over rugged terrain. In this research area, it was shown that the MCD43A3 C6 products with full inversion algorithm, were generally in agreement with the aggregated coarse scale reference albedos over rugged terrain in the Heihe River Basin, with the BSA RMSE of 0.0305 and WSA RMSE of 0.0321, respectively, which were slightly higher than those over flat terrain. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The coarse albedo validation procedure over rugged terrain.</p>
Full article ">Figure 2
<p>The contributions of incident and outgoing radiant flux over a coarse scale pixel; direct, diffuse and terrain irradiances.</p>
Full article ">Figure 3
<p>Simulated DEM with Gaussian height distributions.</p>
Full article ">Figure 4
<p>Study area: (<b>A</b>) the overview of the study area; (<b>B</b>) the location of the Automatic Weather Stations (AWSs) in the DEM imageries; and (<b>C</b>) one example of the AWSs.</p>
Full article ">Figure 5
<p>Scatter plots and the histogram evaluate the aggregated coarse scale albedo (using a Mountain-Radiation-Transfer (MRT)-based model) against the reference simulated coarse scale albedo for (<b>A</b>) the aggregated coarse scale BSA vs. the simulated coarse scale BSA; (<b>B</b>) the bias histogram of the aggregated coarse scale BSA minus the reference coarse scale BSA; (<b>C</b>) the aggregated coarse scale WSA vs. the simulated coarse scale WSA; and (<b>D</b>) the bias between the nine aggregated coarse scale WSAs and the reference simulated WSAs following the increase of the mean slope.</p>
Full article ">Figure 6
<p>Validation results at different solar zenith angle: (<b>A</b>) the solar zenith angle is 0°; (<b>B</b>) the Solar Zenith Angle (SZA) is 20°; (<b>C</b>) the SZA is 40°; and (<b>D</b>) the SZA is 60°.</p>
Full article ">Figure 7
<p>Validation of the MRT-based upscaling model at different slopes (<b>A</b>) the slope is smaller than 10 degree; (<b>B</b>) the slope is greater than 10 degree and less than 20 degree; (<b>C</b>) the slope is between 20 and 30 degrees; and (<b>D</b>) the slope is larger than 30 degrees.</p>
Full article ">Figure 7 Cont.
<p>Validation of the MRT-based upscaling model at different slopes (<b>A</b>) the slope is smaller than 10 degree; (<b>B</b>) the slope is greater than 10 degree and less than 20 degree; (<b>C</b>) the slope is between 20 and 30 degrees; and (<b>D</b>) the slope is larger than 30 degrees.</p>
Full article ">Figure 8
<p>Scatter plots and the histogram evaluate the fine scale albedo against simulated coarse scale albedo for: (<b>A</b>) the aggregated fine scale BSA vs. the simulated coarse scale BSA; (<b>B</b>) the histogram of the aggregated fine scale BSA minus the coarse scale BSA; (<b>C</b>) the aggregated fine scale WSA vs. the simulated coarse scale WSA; and (<b>D</b>) the bias distribution of the aggregated fine scale WSA and the coarse scale WSA following the increase of the mean slopes.</p>
Full article ">Figure 9
<p>Scatter plots and bias histogram between fine scale albedos and the in situ albedos: (<b>A</b>) in situ albedo vs. the fine scale albedo; (<b>B</b>) Bias histogram of fine scale blue-sky albedo minus the in situ albedo The colors refer to the density of points (from highest (red) to lowest (blue).</p>
Full article ">Figure 10
<p>Coarse scale MCD43A3 C6 albedos validation by comparison with the aggregated coarse scale albedos: (<b>A</b>) the aggregated coarse scale BSAs vs. the MCD43A3 C6 BSAs; (<b>B</b>) the aggregated coarse scale WSAs vs. the MCD43A3 C6 WSAs. The colors refer to the density of points (from highest (red) to lowest (blue).</p>
Full article ">Figure 11
<p>Coarse scale MCD43A3 C6 products validation by comparing with the aggregated fine scale albedo at different mean slopes. The colors refer to the density of points (from highest (red) to lowest (blue).</p>
Full article ">Figure 11 Cont.
<p>Coarse scale MCD43A3 C6 products validation by comparing with the aggregated fine scale albedo at different mean slopes. The colors refer to the density of points (from highest (red) to lowest (blue).</p>
Full article ">
16 pages, 3631 KiB  
Article
Design of a Novel Spectral Albedometer for Validating the MODerate Resolution Imaging Spectroradiometer Spectral Albedo Product
by Hongmin Zhou, Jindi Wang and Shunlin Liang
Remote Sens. 2018, 10(1), 101; https://doi.org/10.3390/rs10010101 - 12 Jan 2018
Cited by 7 | Viewed by 4499
Abstract
Land surface shortwave broadband albedo is a key parameter in general circulation models and surface energy budget models. Multispectral satellite data are typically used to generate broadband albedo products in a three-step process: atmospheric correction, for converting the top-of-atmosphere observations to surface directional [...] Read more.
Land surface shortwave broadband albedo is a key parameter in general circulation models and surface energy budget models. Multispectral satellite data are typically used to generate broadband albedo products in a three-step process: atmospheric correction, for converting the top-of-atmosphere observations to surface directional reflectance; angular modeling, for converting the surface directional reflectance to spectral albedo of each individual band; and finally, narrowband-to-broadband conversion, for transforming the spectral albedos to broadband albedos. Spectroradiometers can be used for validating surface directional reflectance products and pyranometers or broadband albedometers, for validating broadband albedo products, but spectral albedo products are rarely validated using ground measurements. In this study, we designed a new type of albedometer that can measure spectral albedos. It consists of multiple interference filters and a silicon detector, for measuring irradiance from 400–1100 nm. The linearity of the sensors is 99%, and the designed albedometer exhibits consistency up to 0.993, with a widely-used commercial instrument. A field experiment for measuring spectral albedo of grassland using this new albedometer was conducted in Yudaokou, China and the measurements are used for validating the MODerate Resolution Imaging Spectroradiometer (MODIS) spectral albedos. The results show that the biases of the MODIS spectral albedos of the first four bands are −0.0094, 0.0065, 0.0159, and −0.0001, respectively. This new instrument provides an effective technique for validating spectral albedos of any satellite sensor in this spectral range, which is critical for improving satellite broadband albedo products. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Spectral albedometer structure. For each band, upwards and downwards sensors are mounted back-to-back, to receive downwards and upwards radiation, respectively.</p>
Full article ">Figure 2
<p>Sensor design. The first layer is a Teflon TM diffuser, the second is an interference filter, the third is a silicon photoelectric detector. Two sensors are included in each band’s albedometer: one for <math display="inline"> <semantics> <mrow> <msub> <mi>I</mi> <mo>↓</mo> </msub> </mrow> </semantics> </math> and one for <math display="inline"> <semantics> <mrow> <msub> <mi>I</mi> <mo>↑</mo> </msub> </mrow> </semantics> </math>.</p>
Full article ">Figure 3
<p>Developed spectral albedometer for one band.</p>
Full article ">Figure 4
<p>Customized interference filter transmittance versus MODIS band response functions. The top figure of each column is the transmittance of the customized interference filter; the bottom figure is the spectral response function of the MODIS band.</p>
Full article ">Figure 5
<p>Spectral albedometer calibration. (<b>a</b>) Incident light is first measured by the spectral albedometer and the radiation of each band is integrated; (<b>b</b>) Incident light is then measured by the spectral albedometer sensor. The upward and downward sensors are calibrated independently.</p>
Full article ">Figure 6
<p>Calibration result of the worst-fitting sensor in the total of 80 calibrated sensors, with coefficient of determination of 0.9996.</p>
Full article ">Figure 7
<p>Non-linear error histogram of 80 calibrated sensors. More than 40% of the sensors are absolutely linear, with a non-linear error of 0; all sensors’ non-linear error is below 0.01.</p>
Full article ">Figure 8
<p>Coefficients of determination of linear fitting results.</p>
Full article ">Figure 9
<p>Spectral albedometer compared with CNR4 radiometer.</p>
Full article ">Figure 10
<p>Field observation of the designed spectral albedometer. The left side shows the tower location from Google Earth on 30 April 2013. The site is homogeneous grassland, except for a road passing through the area. The right figure is the 20-m tower. The designed spectral albedometers are mounted atop the tower.</p>
Full article ">Figure 11
<p>Direct validation of MODIS spectral albedo (bands 1 to 4) with spectral albedometer measurements. The first row are results for 2014 and the second row are results for 2015.</p>
Full article ">
19 pages, 3647 KiB  
Article
Analysis of the Spatial Variability of Land Surface Variables for ET Estimation: Case Study in HiWATER Campaign
by Xiaojun Li, Xiaozhou Xin, Zhiqing Peng, Hailong Zhang, Chuanxiang Yi and Bin Li
Remote Sens. 2018, 10(1), 91; https://doi.org/10.3390/rs10010091 - 11 Jan 2018
Cited by 20 | Viewed by 3822
Abstract
Heterogeneity, including the inhomogeneity of landscapes and surface variables, significantly affects the accuracy of evapotranspiration (ET) (or latent heat flux, LE) estimated from remote sensing satellite data. However, most of the current research uses statistical methods in the mixed pixel to correct the [...] Read more.
Heterogeneity, including the inhomogeneity of landscapes and surface variables, significantly affects the accuracy of evapotranspiration (ET) (or latent heat flux, LE) estimated from remote sensing satellite data. However, most of the current research uses statistical methods in the mixed pixel to correct the ET or LE estimation error, and there is a lack of research from the perspective of the remote sensing model. The method of using frequency distributions or generalized probability density functions (PDFs), which is called the “statistical-dynamical” approach to describe the heterogeneity of land surface characteristics, is a good way to solve the problem. However, in attempting to produce an efficient PDF-based parameterization of remotely sensed ET or LE, first and foremost, it is necessary to systematically understand the variables that are most consistent with the heterogeneity (i.e., variability for a fixed target area or landscape, where the variation in the surface parameter value is primarily concerned with the PDF-based model) of surface turbulence flux. However, the use of PDF alone does not facilitate direct comparisons of the spatial variability of surface variables. To address this issue, the objective of this study is to find an indicator based on PDF to express variability of surface variables. We select the dimensionless or dimensional consistent coefficient of variation (CV), Gini coefficient and entropy to express variability. Based on the analysis of simulated data and field experimental data, we find that entropy is more stable and accurate than the CV and Gini coefficient for expressing the variability of surface variables. In addition, the results of the three methods show that the variability of the leaf area index (LAI) is greater than that of the land surface temperature (LST). Our results provide a suitable method for comparing the variability of different variables. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Geographical information of the study area: (<b>a</b>) schematic of Zhangye city and middle Heihe River Basin; (<b>b</b>) locations of HiWATER-MUSOEXE observation matrix; (<b>c</b>) detailed locations of the 17 EC systems and AWSs inside the Zhangye oasis. EC/AWS is the abbreviation for eddy covariance/automatic weather station.</p>
Full article ">Figure 2
<p>Framework and flowchart of this paper.</p>
Full article ">Figure 3
<p>The PDF distribution of simulated data in Scheme 1. (<b>a</b>–<b>f</b>) refer to Gaussian distribution, beta distribution, gamma distribution, weibull distribution, exponential distribution and uniform distribution, respectively.</p>
Full article ">Figure 4
<p>The PDF distribution of simulated data: (<b>a</b>) Scheme 2; (<b>b</b>) Scheme 3.</p>
Full article ">Figure 5
<p>Data for each region on track 5-1: (<b>a</b>) 1-m classification data derived from CASI and (<b>b</b>) LAI and (<b>c</b>) LST retrieved by WiDAS.</p>
Full article ">Figure 6
<p>Probability density distribution and histogram of the surface variables and the surface types in three study areas: (<b>a</b>) histogram of 1-m classification data; (<b>b</b>) probability density distribution of LAI; (<b>c</b>) frequency distribution histograms of LAI (bandwidth = 1); and (<b>d</b>) probability density distribution of LST. Note: in the classification histogram, the values 1–14 represent maize, leeks, aspen, cauliflower, potato, lettuce, orchard, melon, beam, pear, peper, unclass, shadow and non-veg, respectively.</p>
Full article ">Figure 7
<p>Variation of the three methods when the LST scale is expanded. Note: The abscissa indicates the length of the study area that gradually increases from the upper left corner of <a href="#remotesensing-10-00091-f005" class="html-fig">Figure 5</a>c with a gradient of pixels 5i × 5i (i = 0, 1, 2 … 50).</p>
Full article ">Figure 8
<p>Temporal variation of the entropy and CV of surface variables (i.e., H, LE and Ts_0cm).</p>
Full article ">
20 pages, 10868 KiB  
Article
Comparative Analysis of Chinese HJ-1 CCD, GF-1 WFV and ZY-3 MUX Sensor Data for Leaf Area Index Estimations for Maize
by Jing Zhao, Jing Li, Qinhuo Liu, Hongyan Wang, Chen Chen, Baodong Xu and Shanlong Wu
Remote Sens. 2018, 10(1), 68; https://doi.org/10.3390/rs10010068 - 5 Jan 2018
Cited by 18 | Viewed by 6822
Abstract
In recent years, China has developed and launched several satellites with high spatial resolutions, such as the resources satellite No. 3 (ZY-3) with a multi-spectral camera (MUX) and 5.8 m spatial resolution, the satellite GaoFen No. 1 (GF-1) with a wide field of [...] Read more.
In recent years, China has developed and launched several satellites with high spatial resolutions, such as the resources satellite No. 3 (ZY-3) with a multi-spectral camera (MUX) and 5.8 m spatial resolution, the satellite GaoFen No. 1 (GF-1) with a wide field of view (WFV) camera and 16 m spatial resolution, and the environment satellite (HJ-1A/B) with a charge-coupled device (CCD) sensor and 30 m spatial resolution. First, to analyze the potential application of ZY-3 MUX, GF-1 WFV, and HJ-1 CCD to extract the leaf area index (LAI) at the regional scale, this study estimated LAI from the relationships between physical model-based spectral vegetation indices (SVIs) and LAI values that were generated from look-up tables (LUTs), simulated from the combination of the PROSPECT-5B leaf model and the scattering by arbitrarily inclined leaves with the hot-spot effect (SAILH) canopy reflectance model. Second, to assess the surface reflectance quality of these sensors after data preprocessing, the well-processed surface reflectance products of the Landsat-8 operational land imager (OLI) sensor with a convincing data quality were used to compare the performances of ZY-3 MUX, GF-1 WFV, and HJ-1 CCD sensors both in theory and reality. Apart from several reflectance fluctuations, the reflectance trends were coincident, and the reflectance values of the red and near-infrared (NIR) bands were comparable among these sensors. Finally, to analyze the accuracy of the LAI estimated from ZY-3 MUX, GF-1 WFV, and HJ-1 CCD, the LAI estimations from these sensors were validated based on LAI field measurements in Huailai, Hebei Province, China. The results showed that the performance of the LAI that was inversed from ZY-3 MUX was better than that from GF-1 WFV, and HJ-1 CCD, both of which tended to be systematically underestimated. In addition, the value ranges and accuracies of the LAI inversions both decreased with decreasing spatial resolution. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The geographic location of the study area of ZY-3 MUX based on the false color composite (NIR-red-green) (<b>a</b>), and the subset study area with leaf area index (LAI) field measurements for maize (yellow dots) (<b>b</b>).</p>
Full article ">Figure 2
<p>Spectral response curves for ZY-3 MUX, GF-1 WFV, and HJ-1 CCD from CRESDA, and for Landsat-8 OLI from USGS.</p>
Full article ">Figure 3
<p>Flowchart of the LAI inversion method based on ZY-3 MUX, GF-1 WFV, and HJ-1 CCD.</p>
Full article ">Figure 4
<p>LAI inversion results for ZY-3 MUX (<b>a</b>,<b>d</b>), GF-1 WFV (<b>b</b>,<b>e</b>), and HJ-1 CCD (<b>c</b>,<b>f</b>) in Huailai, Hebei Province.</p>
Full article ">Figure 5
<p>Comparisons of the LAI inversions with the field measurements for ZY-3 MUX (<b>a</b>,<b>d</b>), GF-1 WFV(<b>b</b>,<b>e</b>), and HJ-1 CCD (<b>c</b>,<b>f</b>).</p>
Full article ">Figure 6
<p>Validation of the LAI inversions from ZY-3 MUX, GF-1 WFV, and HJ-1 CCD using field data with high spatial representativeness (<b>a</b>) and low spatial representativeness (<b>b</b>).</p>
Full article ">Figure 7
<p>LAI inversions up-scaled to 30 m spatial resolution from ZY-3 MUX (<b>a</b>), GF-1 WFV (<b>b</b>), and HJ-1 CCD (<b>c</b>).</p>
Full article ">Figure 8
<p>LAI inversions up-scaled to 16 m spatial resolution from ZY-3 MUX plotted against GF-1 WFV LAI inversions (<b>a</b>) and LAI inversions up-scaled to 30 m spatial resolution from ZY-3 MUX (<b>b</b>) or GF-1 WFV (<b>c</b>) plotted against HJ-1 CCD LAI inversions for pure (blue dots) and mixed (red dots) pixels.</p>
Full article ">Figure 9
<p>Reflectance in the red band (<b>a</b>) and the NIR band (<b>b</b>) varied with LAI field measurements for ZY-3 MUX, GF-1 WFV, and HJ-1 CCD.</p>
Full article ">Figure 10
<p>Reflectance in the red and NIR bands with LAI from 0 to 8 at 0.1 intervals for ZY-3 MUX, GF-1 WFV, HJ-1 CCD, and Landsat-8 OLI (<b>a</b>), and difference in reflectance in the red (<b>b</b>) and NIR (<b>c</b>) bands between Landsat-8 OLI and each of ZY-3 MUX, GF-1 WFV, and HJ-1 CCD for maize simulations.</p>
Full article ">Figure 11
<p>Density scatter plots of the reflectance difference in the red and NIR bands between Landsat-8 OLI and each of ZY-3 MUX (<b>a</b>), GF-1 WFV (<b>b</b>), and HJ-1 CCD (<b>c</b>) for croplands.</p>
Full article ">
6727 KiB  
Article
High-Resolution Mapping of Freeze/Thaw Status in China via Fusion of MODIS and AMSR2 Data
by Tongxi Hu, Tianjie Zhao, Jiancheng Shi, Shengli Wu, Dan Liu, Haiming Qin and Kaiguang Zhao
Remote Sens. 2017, 9(12), 1339; https://doi.org/10.3390/rs9121339 - 20 Dec 2017
Cited by 18 | Viewed by 4983
Abstract
Transition of freeze/thaw (F/T) affects land-atmospheric interactions and other biospheric dynamics. Global F/T statuses are normally monitored using microwave remote sensing, but at coarse resolutions (e.g., 25 km). Integration of coarse microwave remote sensing data with finer satellite products represents an opportunity to [...] Read more.
Transition of freeze/thaw (F/T) affects land-atmospheric interactions and other biospheric dynamics. Global F/T statuses are normally monitored using microwave remote sensing, but at coarse resolutions (e.g., 25 km). Integration of coarse microwave remote sensing data with finer satellite products represents an opportunity to further enhance our ability to map F/T statuses regionally and globally. Here, we implemented and tested an approach to generate daily F/T status maps at a 5-km spatial resolution through the fusion of passive microwave data from AMSR2 and land surface temperature products from MODIS, using China as our study area for the year 2013 and 2014. Moreover, possible influences from elevation, vegetation, seasonality, etc., were also analyzed, as such analysis provides a direction to improve the approach. Overall, our freeze/thaw maps agreed well with ground reference observations, with an accuracy of ~86.6%. The new F/T maps helped to identify regions subject to frequent F/T transitions through the year, such as the Qinghai-Tibetan Plateau, Xinjiang, Gansu, Heilongjiang, Jilin, and Liaoning Province. This study indicates that the combination of AMSR2 and MODIS observations provides an effective method to obtain finer F/T maps (5-km or lower) for extensive regions. The finer F/T maps improve our knowledge of the F/T state detected by satellite remote sensing, and have a wide range of applications in regional studies considering land surface heterogeneity and models (e.g., community land models). Full article
Show Figures


<p>Parameter “a” and NDVI on forest (<b>a</b>), barren (<b>b</b>), grass land (<b>c</b>) and crop land (<b>d</b>). P_a_asc and P_a_des are for observations at 13:30 and 1:30, respectively.</p>
Full article ">
<p>Parameter “a” and NDVI on forest (<b>a</b>), barren (<b>b</b>), grass land (<b>c</b>) and crop land (<b>d</b>). P_a_asc and P_a_des are for observations at 13:30 and 1:30, respectively.</p>
Full article ">Graphical abstract
Full article ">Figure 1
<p>Flowchart of generating high-resolution F/T maps.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>b</b>) are correlations of the MODIS LST and freeze/thaw index (FTI) for observations at 13:30 and 1:30, respectively.</p>
Full article ">Figure 3
<p>Spatial distribution of parameters “a” and “b” in regression models for observations at 13:30 and 1:30, respectively. (<b>a</b>,<b>b</b>) are for “a”; (<b>c</b>,<b>d</b>) are for “b”.</p>
Full article ">Figure 4
<p>Coverage of MODIS observations over China in 2013 (<b>a</b>) and frequency of valid MODIS observations for each pixel at 13:30 (<b>b</b>) and 1:30 (<b>c</b>) in 2013.</p>
Full article ">Figure 5
<p>Temporal agreement between in situ measurement and F/T maps for observations at 13:30 (<b>a</b>) and 1:30 (<b>b</b>).</p>
Full article ">Figure 6
<p>Accuracy assessment of F/T maps at individual meteorological stations in China.</p>
Full article ">Figure 7
<p>Correlation of the MODIS LST and FTI derived by microwave observations from 70 randomly-selected stations at different elevations. Ele is the elevation of the stations; Coeff_a is the correaltion at 13:30; Coeff_d is the correlation at 1:30.</p>
Full article ">Figure 8
<p>Correlation of the MODIS LST and FTI derived by microwave observations from 70 randomly-selected stations with different annual average NDVI. Corr_a is the correaltion at 13:30; Corr_d is the correlation at 1:30.</p>
Full article ">
7914 KiB  
Article
MODIS-Based Estimation of Terrestrial Latent Heat Flux over North America Using Three Machine Learning Algorithms
by Xuanyu Wang, Yunjun Yao, Shaohua Zhao, Kun Jia, Xiaotong Zhang, Yuhu Zhang, Lilin Zhang, Jia Xu and Xiaowei Chen
Remote Sens. 2017, 9(12), 1326; https://doi.org/10.3390/rs9121326 - 16 Dec 2017
Cited by 26 | Viewed by 6564
Abstract
Terrestrial latent heat flux (LE) is a key component of the global terrestrial water, energy, and carbon exchanges. Accurate estimation of LE from moderate resolution imaging spectroradiometer (MODIS) data remains a major challenge. In this study, we estimated the daily LE for different [...] Read more.
Terrestrial latent heat flux (LE) is a key component of the global terrestrial water, energy, and carbon exchanges. Accurate estimation of LE from moderate resolution imaging spectroradiometer (MODIS) data remains a major challenge. In this study, we estimated the daily LE for different plant functional types (PFTs) across North America using three machine learning algorithms: artificial neural network (ANN); support vector machines (SVM); and, multivariate adaptive regression spline (MARS) driven by MODIS and Modern Era Retrospective Analysis for Research and Applications (MERRA) meteorology data. These three predictive algorithms, which were trained and validated using observed LE over the period 2000–2007, all proved to be accurate. However, ANN outperformed the other two algorithms for the majority of the tested configurations for most PFTs and was the only method that arrived at 80% precision for LE estimation. We also applied three machine learning algorithms for MODIS data and MERRA meteorology to map the average annual terrestrial LE of North America during 2002–2004 using a spatial resolution of 0.05°, which proved to be useful for estimating the long-term LE over North America. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Architecture of the neural network model used in this study.</p>
Full article ">Figure 2
<p>One-dimensional linear regression with <math display="inline"> <semantics> <mi>ε</mi> </semantics> </math>-insensitive band for the support vector machine (SVM) algorithm.</p>
Full article ">Figure 3
<p>The hinge functions and knot location in the multivariate adaptive regression spline (MARS) model.</p>
Full article ">Figure 4
<p>Location of the 85 eddy covariance flux towers used in this study. INV means the data of this site were used to inverse, TRA means that the data of this site were used for training.</p>
Full article ">Figure 5
<p>Bar graphs of the training and validation statistics (<span class="html-italic">R</span><sup>2</sup>, Bias and root mean square error (RMSE)) of three algorithms driven by tower-specific meteorology for five PFTs at the 85 flux tower site. All <span class="html-italic">R</span><sup>2</sup> values are significant with a 99% confidence.</p>
Full article ">Figure 6
<p>Examples of the eight-day terrestrial latent heat flux (LE) average as measured and estimated using different machine learning algorithms for the different PFTs.</p>
Full article ">Figure 6 Cont.
<p>Examples of the eight-day terrestrial latent heat flux (LE) average as measured and estimated using different machine learning algorithms for the different PFTs.</p>
Full article ">Figure 7
<p>Bar graphs of the training and validation statistics (<span class="html-italic">R</span><sup>2</sup>, Bias and RMSE) of three algorithms driven by MERRA meteorology for five PFTs at the 85 flux tower sites. All of the <span class="html-italic">R</span><sup>2</sup> values are significant with a 99% confidence.</p>
Full article ">Figure 8
<p>The map of mean annual terrestrial LE from 2002 to 2004 at a spatial resolution of 0.05° using three machine learning algorithms driven by MERRA meteorology over North America.</p>
Full article ">Figure 9
<p>Comparison of daily LE observations for all 85 flux tower sites and LE estimates using different machine learning algorithms driven by tower-specific meteorology.</p>
Full article ">Figure 10
<p>Comparison of daily LE observations for all 85 flux tower sites and LE estimates using different machine learning algorithms driven by MERRA meteorology.</p>
Full article ">Figure 11
<p>Spatial differences in the average annual terrestrial LE (2002–2004) between three machine learning algorithms. .</p>
Full article ">Figure 12
<p>Spatial differences in the average annual terrestrial LE (2002–2004) between MODIS LE product and LE product using three machine learning algorithms.</p>
Full article ">
4124 KiB  
Article
Detecting Forest Disturbance in Northeast China from GLASS LAI Time Series Data Using a Dynamic Model
by Jian Wang, Jindi Wang, Hongmin Zhou and Zhiqiang Xiao
Remote Sens. 2017, 9(12), 1293; https://doi.org/10.3390/rs9121293 - 12 Dec 2017
Cited by 14 | Viewed by 4669
Abstract
Large-scale forest disturbance often leads to changes in forest cover and structure, which imposes a great uncertainty in the estimation of the forest carbon cycle and biomass and affects other applications. In northeastern China, the Daxinganling region has abundant forest resources, where the [...] Read more.
Large-scale forest disturbance often leads to changes in forest cover and structure, which imposes a great uncertainty in the estimation of the forest carbon cycle and biomass and affects other applications. In northeastern China, the Daxinganling region has abundant forest resources, where the forest coverage is about 30%. The Global LAnd Surface Satellite (GLASS) leaf area index (LAI) time series data provide important information to monitor the possible change of forests. In this study, we developed a new method to detect forest disturbances using GLASS LAI data over the Daxinganling region of Northeast China. As a dynamic model, the season-trend model has a higher sensitivity toward a seasonal change in LAI. Based on the accumulation of multi-year GLASS LAI products from 1997 to 2002, the dynamic model of LAI time series for each pixel is established first. The time-stepping modeling (TSM) process was designed by using the season-trend method, and sequential tests for detecting disturbances from a time series of pixels. Significant changes in the model parameters were captured as disturbance signals. Then, the near-infrared and shortwave-infrared bands of Moderate Resolution Imaging Spectroradiometer (MODIS) surface reflectance are used as auxiliary information to distinguish the types of forest disturbances. Here, the algorithm led to the detection of two different types of disturbances: fire and other (e.g., insect, drought, deforestation). In this study, we took the forest region as the study area, used the 8-day composite GLASS LAI data at 1000-m spatial resolution to identify each pixel as a fire disturbance, other disturbance, or non-disturbance. Validation was performed using reference burned area data derived from Landsat 30 m imagery. Results were also compared with the MCD64 product. The validation results were based on confusion matrices showing the overall accuracy (OA) exceeded 92% for our method and the MCD64 product. Statistical tests identified that TSM’s product accuracy is higher than that of MCD64. This study demonstrated that the TSM algorithm using a season-trend model provides a simple and automated approach to identify and map forest disturbance. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Map showing the location of the study area. (<b>A</b>) Land cover map of study area in 2003 produced by Landsat images of 30 m resolution; (<b>B</b>) Land cover mapping with 1000 m resolution.</p>
Full article ">Figure 2
<p>Flow chart of the forest disturbance detection. LAI = leaf area index; DI = disturbance index; NBR = normalized burn ratio.</p>
Full article ">Figure 3
<p>Using the season-trend model to simulate GLASS LAI time series compared to the previous two years LAI values in 2003. (<b>A</b>) multiplied the coefficient 4/3, (<b>B</b>) is 1, (<b>C</b>) is 2/3, and (<b>D</b>) is 1/3. Where the red line is the curve of the amplitude of the parameter c, the purple line is the amplitude curve of b.</p>
Full article ">Figure 4
<p>Difference between the NBR anomalies for burned (<b>A</b>) and unburned (<b>B</b>) areas in 2003. The green curve is the value of NBR, and the red dotted line is the horizontal line when NBR is zero.</p>
Full article ">Figure 5
<p>Comparison of GLASS burned area data and the MCD64A1 product data.</p>
Full article ">Figure 6
<p>The LAI (green) and NBR (red) time series in the study area. (<b>A</b>,<b>B</b>) are detected as burned pixels in the TSM algorithm, and (<b>C</b>,<b>D</b>) are detected as unburned.</p>
Full article ">Figure 7
<p>Other types of forest disturbances and mean LAI (subscript is 1) and NBR (subscript is 2) time series of two verification regions. Region <b>A</b>: The maximum mean LAI in 2003 at a value of 2.1; maximum mean LAI was 4.1 for 2000 and 2001. Region <b>B</b>: The maximum mean LAI in 2003 at a value of 2.8; maximum mean LAI was 4.5 for 2000 and 2001.</p>
Full article ">Figure 8
<p>Regional images of the 2003 GLASS and Landsat fire disturbance in Shibazhan and Jinhe. The comparison results of two algorithms in Shibazhan are shown in the upper row: (<b>a</b>) TSM results; (<b>b</b>) Landsat image; (<b>c</b>) Landsat dNBR results. The comparison results of two algorithms in Jinhe are shown in the bottom row: (<b>d</b>) TSM results; (<b>e</b>) Landsat image; (<b>f</b>) Landsat dNBR results. The Landsat image is displayed in band 4 (red), 3 (green), 2 (blue) color composite. The cyan boxes in the differenced NBR (dNBR) represent 10 × 10 GLASS pixels.</p>
Full article ">Figure 9
<p>The time series curve of different products at the same location (3 × 3 km); GLASS LAI (red), MODIS LAI (green) and MODIS normalized difference vegetation index (NDVI) (black).</p>
Full article ">
6241 KiB  
Article
Uncertainty of Remote Sensing Data in Monitoring Vegetation Phenology: A Comparison of MODIS C5 and C6 Vegetation Index Products on the Tibetan Plateau
by Zhoutao Zheng and Wenquan Zhu
Remote Sens. 2017, 9(12), 1288; https://doi.org/10.3390/rs9121288 - 11 Dec 2017
Cited by 26 | Viewed by 5135
Abstract
Vegetation phenology is considered a sensitive indicator of climate change, which controls carbon, nitrogen, and water cycles within terrestrial ecosystems. The Moderate Resolution Imaging Spectroradiometer (MODIS) Normalized Difference Vegetation Index (NDVI) is an important moderate resolution remote sensing data for monitoring vegetation phenology. [...] Read more.
Vegetation phenology is considered a sensitive indicator of climate change, which controls carbon, nitrogen, and water cycles within terrestrial ecosystems. The Moderate Resolution Imaging Spectroradiometer (MODIS) Normalized Difference Vegetation Index (NDVI) is an important moderate resolution remote sensing data for monitoring vegetation phenology. However, Terra MODIS Collection 5 (C5) vegetation index products were identified to be affected by sensor degradation, which has been addressed in the recently released MODIS Collection 6 (C6) vegetation index products. In order to compare the difference between MODIS C5 and C6 NDVI in monitoring vegetation phenology, the start and end of growing season (SOS and EOS) of the alpine grassland on the Tibetan Plateau (TP) were extracted using four common methods. Then, the C5 and C6 NDVI-derived SOS (SOSC5 and SOSC6) and EOS (EOSC5 and EOSC6) were compared with ground-observed phenology data. Results showed that the multi-year average growing season NDVIs of C6 were lower than those of C5 in most areas, while the inter-annual variation patterns of regional average SOSC5 and SOSC6 (EOSC5 and EOSC6) were consistent. However, large spatial differences in phenological trends were found between C5 and C6 NDVI products. From C5 to C6, pixels with a SOS (EOS) trend shifting from significant to insignificant or from insignificant to significant accounted for at least 14.58% (9.07%) of the total pixels. SOSC5 was more consistent than SOSC6 with the ground-observed green-up dates. C5 NDVI may be more appropriate for monitoring SOS than C6 NDVI in the study region, but more ground-observed phenology records are needed to confirm it due to only four observational sites in this study. However, large differences and poor correlations existed between EOSC5 (EOSC6) and the ground-observed beginning of leaf coloring. To further evaluate the uncertainty of MODIS C5 and C6 NDVI in monitoring vegetation phenology, higher resolution near-surface remote sensing data and corresponding validation methods should be applied. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Comparison of annual regional average growing season Normalized Difference Vegetation Index derived from MODIS C5 (GSNDVI<sub>C5</sub>) and C6 (GSNDVI<sub>C6</sub>) for the alpine grassland on the Tibetan Plateau during 2001–2015.</p>
Full article ">Figure 2
<p>(<b>a</b>) The relative difference in the multi-year average growing season Normalized Difference Vegetation Index between C6 (GSNDVI<sub>C6</sub>) and C5 (GSNDVI<sub>C5</sub>) ((C6–C5)/C5) during 2001–2015; (<b>b</b>) The scatter plot of the multi-year average GSNDVI<sub>C6</sub> against GSNDVI<sub>C</sub><sub>5</sub> during 2001–2015.</p>
Full article ">Figure 3
<p>(<b>a</b>) Comparison of trends in growing season Normalized Difference Vegetation Index between C5 (GSNDVI<sub>C5</sub>) and C6 (GSNDVI<sub>C6</sub>) during 2001–2015; and, (<b>b</b>) The histograms of trends in GSNDVI<sub>C5</sub> and GSNDVI<sub>C6</sub> during 2001–2015.</p>
Full article ">Figure 4
<p>Comparisons of annual regional average start of growing season from MODIS C5 (SOS<sub>C5</sub>) and C6 (SOS<sub>C6</sub>) for the alpine grassland on the Tibetan Plateau during 2001–2015 for different methods: (<b>a</b>) Maximum Curvature Change (MCC); (<b>b</b>) Dynamic Threshold 0.2 (DT2); (<b>c</b>) Dynamic Threshold 0.5 (DT5); and, (<b>d</b>) Maximum Slope (MS).</p>
Full article ">Figure 5
<p>Comparisons of annual regional average end of growing season from MODIS C5 (SOS<sub>C5</sub>) and C6 (SOS<sub>C6</sub>) for the alpine grassland on the Tibetan Plateau during 2001–2015 for different methods: (<b>a</b>) Maximum Curvature Change (MCC); (<b>b</b>) Dynamic Threshold 0.2 (DT2); (<b>c</b>) Dynamic Threshold 0.5 (DT5); (<b>d</b>) Maximum Slope (MS).</p>
Full article ">Figure 6
<p>Spatial differences between the multi-year average SOS<sub>C6</sub> and SOS<sub>C5</sub> (C6–C5) for different methods: (<b>a</b>) Maximum Curvature Change (MCC); (<b>b</b>) Dynamic Threshold 0.2 (DT2); (<b>c</b>) Dynamic Threshold 0.5 (DT5); (<b>d</b>) Maximum Slope (MS).</p>
Full article ">Figure 7
<p>Spatial differences between the multi-year average EOS<sub>C6</sub> and EOS<sub>C5</sub> (C6–C5) for different methods: (<b>a</b>) Maximum Curvature Change (MCC); (<b>b</b>) Dynamic Threshold 0.2 (DT2); (<b>c</b>) Dynamic Threshold 0.5 (DT5); and, (<b>d</b>) Maximum Slope (MS).</p>
Full article ">Figure 8
<p>Comparison of trends in start of growing season between C5 (SOS<sub>C5</sub>) and C6 (SOS<sub>C6</sub>) during 2001–2015 for different methods: (<b>a</b>) Maximum Curvature Change (MCC); (<b>b</b>) Dynamic Threshold 0.2 (DT2); (<b>c</b>) Dynamic Threshold 0.5 (DT5); and, (<b>d</b>) Maximum Slope (MS).</p>
Full article ">Figure 9
<p>Comparison of trends in end of growing season between C5 (EOS<sub>C5</sub>) and C6 (EOS<sub>C6</sub>) during 2001–2015 for different methods: (<b>a</b>) Maximum Curvature Change (MCC); (<b>b</b>) Dynamic Threshold 0.2 (DT2); (<b>c</b>) Dynamic Threshold 0.5 (DT5); and, (<b>d</b>) Maximum Slope (MS).</p>
Full article ">
5897 KiB  
Article
Estimating Diurnal Courses of Gross Primary Production for Maize: A Comparison of Sun-Induced Chlorophyll Fluorescence, Light-Use Efficiency and Process-Based Models
by Tianxiang Cui, Rui Sun, Chen Qiao, Qiang Zhang, Tao Yu, Gang Liu and Zhigang Liu
Remote Sens. 2017, 9(12), 1267; https://doi.org/10.3390/rs9121267 - 7 Dec 2017
Cited by 17 | Viewed by 5568
Abstract
Accurately quantifying gross primary production (GPP) is of vital importance to understanding the global carbon cycle. Light-use efficiency (LUE) models and process-based models have been widely used to estimate GPP at different spatial and temporal scales. However, large uncertainties remain in quantifying GPP, [...] Read more.
Accurately quantifying gross primary production (GPP) is of vital importance to understanding the global carbon cycle. Light-use efficiency (LUE) models and process-based models have been widely used to estimate GPP at different spatial and temporal scales. However, large uncertainties remain in quantifying GPP, especially for croplands. Recently, remote measurements of solar-induced chlorophyll fluorescence (SIF) have provided a new perspective to assess actual levels of plant photosynthesis. In the presented study, we evaluated the performance of three approaches, including the LUE-based multi-source data synergized quantitative (MuSyQ) GPP algorithm, the process-based boreal ecosystem productivity simulator (BEPS) model, and the SIF-based statistical model, in estimating the diurnal courses of GPP at a maize site in Zhangye, China. A field campaign was conducted to acquire synchronous far-red SIF (SIF760) observations and flux tower-based GPP measurements. Our results showed that both SIF760 and GPP were linearly correlated with APAR, and the SIF760-GPP relationship was adequately characterized using a linear function. The evaluation of the modeled GPP against the GPP measured from the tower demonstrated that all three approaches provided reasonable estimates, with R2 values of 0.702, 0.867, and 0.667 and RMSE values of 0.247, 0.153, and 0.236 mg m−2 s−1 for the MuSyQ-GPP, BEPS and SIF models, respectively. This study indicated that the BEPS model simulated the GPP best due to its efficiency in describing the underlying physiological processes of sunlit and shaded leaves. The MuSyQ-GPP model was limited by its simplification of some critical ecological processes and its weakness in characterizing the contribution of shaded leaves. The SIF760-based model demonstrated a relatively limited accuracy but showed its potential in modeling GPP without dependency on climate inputs in short-term studies. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of the study site.</p>
Full article ">Figure 2
<p>Overview of the spectral measurements: (<b>a</b>) spectral measurement system; and (<b>b</b>) positions of the standard reflectance panels (S) and the four canopy targets (T1–T4).</p>
Full article ">Figure 3
<p>Diurnal patterns of PAR, GPP and SIF<sub>760</sub> during the experiment: (<b>a</b>) 10 July; (<b>b</b>) 17 July; (<b>c</b>) 18 July; (<b>d</b>) 21 August; and (<b>e</b>) 22 August.</p>
Full article ">Figure 4
<p>Relationship between APAR (absorbed photosynthetically active radiation) and: (<b>a</b>) individual SIF<sub>760</sub>; (<b>b</b>) GPP; and (<b>c</b>) averaged SIF. The error bar indicates the range of SIF<sub>760</sub> values for four measurements.</p>
Full article ">Figure 5
<p>Relationship between: (<b>a</b>) individual SIF<sub>760</sub> and GPP; and (<b>b</b>) averaged SIF<sub>760</sub> and GPP. The error bars indicate the range of SIF<sub>760</sub> values for the four measurements.</p>
Full article ">Figure 6
<p>Relationships between modeled GPP and EC-based GPP by: (<b>a</b>) SIF<sub>760</sub>; (<b>b</b>) MuSyQ-GPP; (<b>c</b>) BEPS; and (<b>d</b>) SIF<sub>686</sub> during the experiment period. The dots in the red circle are the values at 8:30 on 21 August, and 8:30 on 22 August. The error bars indicate the range of the four SIF-based GPP values, the red shades represent the 95% confidence bands for the regression functions.</p>
Full article ">Figure 7
<p>Incident radiance at: (<b>a</b>) 8:30 on 21 August; and (<b>b</b>) 8:30 on 22 August. T1 to T4 represent the four canopy targets, B and A denote measurements before and after canopy measurement, respectively.</p>
Full article ">Figure 8
<p><span class="html-italic">DQ</span><sub>s</sub> values of the measurements: (<b>a</b>) 10 July; (<b>b</b>) 17 July; (<b>c</b>) 18 July; (<b>d</b>) 21 August; and (<b>e</b>) 22 August. T1 to T4 represent the measurements of the four canopy targets.</p>
Full article ">
10192 KiB  
Article
Estimating Land Surface Temperature from Feng Yun-3C/MERSI Data Using a New Land Surface Emissivity Scheme
by Xiangchen Meng, Jie Cheng and Shunlin Liang
Remote Sens. 2017, 9(12), 1247; https://doi.org/10.3390/rs9121247 - 1 Dec 2017
Cited by 41 | Viewed by 6083
Abstract
Land surface temperature (LST) is a key parameter for a wide number of applications, including hydrology, meteorology and surface energy balance. In this study, we first proposed a new land surface emissivity (LSE) scheme, including a lookup table-based method to determine the vegetated [...] Read more.
Land surface temperature (LST) is a key parameter for a wide number of applications, including hydrology, meteorology and surface energy balance. In this study, we first proposed a new land surface emissivity (LSE) scheme, including a lookup table-based method to determine the vegetated surface emissivity and an empirical method to derive the bare soil emissivity from the Global LAnd Surface Satellite (GLASS) broadband emissivity (BBE) product. Then, the Modern Era Retrospective-Analysis for Research and Applications (MERRA) reanalysis data and the Feng Yun-3C/Medium Resolution Spectral Imager (FY-3C/MERSI) precipitable water vapor product were used to correct the atmospheric effects. After resolving the land surface emissivity and atmospheric effects, the LST was derived in a straightforward manner from the FY-3C/MERSI data by the radiative transfer equation algorithm and the generalized single-channel algorithm. The mean difference between the derived LSE and field-measured LSE over seven stations is approximately 0.002. Validation of the LST retrieved with the LSE determined by the new scheme can achieve an acceptable accuracy. The absolute biases are less than 1 K and the STDs (RMSEs) are less than 1.95 K (2.2 K) for both the 1000 m and 250 m spatial resolutions. The LST accuracy is superior to that retrieved with the LSE determined by the commonly used Normalized Difference Vegetation Index (NDVI) threshold method. Thus, the new emissivity scheme can be used to improve the accuracy of the LSE and further the LST for sensors with broad spectral ranges such as FY-3C/MERSI. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Spatial distribution of the six in situ sites. (The base map is from HJ-1 CCD false color composite image and the RGB components are channels 4, 3 and 2, respectively).</p>
Full article ">Figure 2
<p>Plot of the atmospheric transmittance (<b>a</b>) or upward radiance (<b>b</b>) at a given VZA against the atmospheric transmittance or upward radiance at nadir view.</p>
Full article ">Figure 3
<p>Spectral response for band 5 of FY-3C/MERSI.</p>
Full article ">Figure 4
<p>Photos of measurement sites. (<b>a</b>) CJZ01; (<b>b</b>) CJZ02; (<b>c</b>) GB01; (<b>d</b>) SSW01; (<b>e</b>) SSW02,03; (<b>f</b>) GB02.</p>
Full article ">Figure 5
<p>LSE simulated by 4SAIL model varying LAI at a fixed soil emissivity of 0.96 (<b>left</b>) or a fixed leaf emissivity of 0.96 (<b>right</b>).</p>
Full article ">Figure 6
<p>The BBE of each eight-day period from 2001 to 2014 for six stations.</p>
Full article ">Figure 7
<p>Scatterplots between estimated LST from RTE algorithm (<b>left</b>) or GSC algorithm (<b>right</b>) and in situ LST.</p>
Full article ">Figure 8
<p>The images and histograms of LSE difference between the AST_05 product and the new method on 17 and 26 July 2014; the corresponding images and histograms of LST difference. (<b>a</b>) The image of LSE difference on 17 July 2014; (<b>b</b>) The image of LSE difference on 26 July 2014; (<b>c</b>) The histogram of LSE difference on 17 July 2014; (<b>d</b>) The histogram of LSE difference on 26 July 2014; (<b>e</b>) The image of LST difference on 17 July 2014; (<b>f</b>) The image of LST difference on 26 July 2014; (<b>g</b>) The histogram of LST difference on 17 July 2014; (<b>h</b>) The histogram of LST difference on 26 July 2014.</p>
Full article ">Figure 9
<p>Boxplots between estimated LSTs from RTE algorithm (<b>left</b>) or GSC algorithm (<b>right</b>) at 250 m resolution and in situ LSTs.</p>
Full article ">
1330 KiB  
Article
Spatiotemporally Representative and Cost-Efficient Sampling Design for Validation Activities in Wanglang Experimental Site
by Gaofei Yin, Ainong Li and Aleixandre Verger
Remote Sens. 2017, 9(12), 1217; https://doi.org/10.3390/rs9121217 - 26 Nov 2017
Cited by 10 | Viewed by 4559
Abstract
Spatiotemporally representative Elementary Sampling Units (ESUs) are required for capturing the temporal variations in surface spatial heterogeneity through field measurements. Since inaccessibility often coexists with heterogeneity, a cost-efficient sampling design is mandatory. We proposed a sampling strategy to generate spatiotemporally representative and cost-efficient [...] Read more.
Spatiotemporally representative Elementary Sampling Units (ESUs) are required for capturing the temporal variations in surface spatial heterogeneity through field measurements. Since inaccessibility often coexists with heterogeneity, a cost-efficient sampling design is mandatory. We proposed a sampling strategy to generate spatiotemporally representative and cost-efficient ESUs based on the conditioned Latin hypercube sampling scheme. The proposed strategy was constrained by multi-temporal Normalized Difference Vegetation Index (NDVI) imagery, and the ESUs were limited within a sampling feasible region established based on accessibility criteria. A novel criterion based on the Overlapping Area (OA) between the NDVI frequency distribution histogram from the sampled ESUs and that from the entire study area was used to assess the sampling efficiency. A case study in Wanglang National Nature Reserve in China showed that the proposed strategy improves the spatiotemporally representativeness of sampling (mean annual OA = 74.7%) compared to the single-temporally constrained (OA = 68.7%) and the random sampling (OA = 63.1%) strategies. The introduction of the feasible region constraint significantly reduces in-situ labour-intensive characterization necessities at expenses of about 9% loss in the spatiotemporal representativeness of the sampling. Our study will support the validation activities in Wanglang experimental site providing a benchmark for locating the nodes of automatic observation systems (e.g., LAINet) which need a spatially distributed and temporally fixed sampling design. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Map of China with the location of Wanglang experimental site. (<b>b</b>) A Landsat 8 Operational Land Imager (OLI) image and (<b>c</b>) an elevation map of the study area. The roads and the feasible accessible region are shown.</p>
Full article ">Figure 2
<p>Overlapping area (OA) of the NDVI histogram distributions for the sampled and entire study area as a function of the number of Elementary Sampling Units (ESU).</p>
Full article ">Figure 3
<p>The spatial distribution of the 20 elementary sampling units for the multi-temporal constrained sampling design in the feasible region.</p>
Full article ">Figure 4
<p>NDVI frequency distribution histograms of 20 selected elementary sampling units for the four sampling strategies (from top to bottom): random sampling (RS), single-temporally constrained sampling (SC), and multi-temporally constrained sampling (MC) in the feasible region (MC) and in the entire study area without the sampling feasible constraint (MCE). The red lines represent the NDVI frequency distribution histograms of the entire study area.</p>
Full article ">Figure 5
<p>Scatterplots between the average sampled NDVI and the average true NDVI values for the entire study area and for the five different dates. (<b>a</b>) Random sampling in the feasible region. (<b>b</b>) Single-temporally constrained sampling in the feasible region. (<b>c</b>) Multi-temporally constrained sampling in the feasible region. (<b>d</b>) Multi-temporally constrained sampling in the entire region. The bars indicate the standard deviations.</p>
Full article ">Figure 6
<p>Density scatterplots between the NDVI and the cosine of the local solar incidence angle (cos(<span class="html-italic">i</span>)) in Wanglang study area for the different acquisition dates.</p>
Full article ">
6103 KiB  
Article
New Scheme for Validating Remote-Sensing Land Surface Temperature Products with Station Observations
by Wenping Yu, Mingguo Ma, Zhaoliang Li, Junlei Tan and Adan Wu
Remote Sens. 2017, 9(12), 1210; https://doi.org/10.3390/rs9121210 - 24 Nov 2017
Cited by 24 | Viewed by 5479
Abstract
Continuous land-surface temperature (LST) observations from ground-based stations are an important reference dataset for validating remote-sensing LST products. However, a lack of evaluations of the representativeness of station observations limits the reliability of validation results. In this study, a new practical validation scheme [...] Read more.
Continuous land-surface temperature (LST) observations from ground-based stations are an important reference dataset for validating remote-sensing LST products. However, a lack of evaluations of the representativeness of station observations limits the reliability of validation results. In this study, a new practical validation scheme is presented for validating remote-sensing LST products that includes a key step: assessing the spatial representativeness of ground-based LST measurements. Three indicators, namely, the dominant land-cover type (DLCT), relative bias (RB), and average structure scale (ASS), are established to quantify the representative levels of station observations based on the land-cover type (LCT) and LST reference maps with high spatial resolution. We validated MODIS LSTs using station observations from the Heihe River Basin (HRB) in China. The spatial representative evaluation steps show that the representativeness of observations greatly differs among stations and varies with different vegetation growth and other factors. Large differences in the validation results occur when using different representative level observations, which indicates a large potential for large error during the traditional T-based validation scheme. Comparisons show that the new validation scheme greatly improves the reliability of LST product validation through high-level representative observations. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>New scheme for land surface temperature (LST) validation based on the assessment of local spatial representativeness (site level). In the scheme, LCT is the abbreviation of land-cover type, and NDVI is the abbreviation of normalized difference vegetation index.</p>
Full article ">Figure 2
<p>Study area and locations of the validation stations.</p>
Full article ">Figure 3
<p>Photos of the 18 Heihe Watershed Allied Telemetry Experimental Research (HiWATER) stations. The two pyrgeometers that were used to record the longwave radiation were deployed at an average height of 6 m, with one facing upwards and the other facing downwards.</p>
Full article ">Figure 3 Cont.
<p>Photos of the 18 Heihe Watershed Allied Telemetry Experimental Research (HiWATER) stations. The two pyrgeometers that were used to record the longwave radiation were deployed at an average height of 6 m, with one facing upwards and the other facing downwards.</p>
Full article ">Figure 4
<p>The dominant land cover type (DLCT) for the HiWATER stations.</p>
Full article ">Figure 5
<p>Monthly relative bias (RB) of the observations from the 18 stations within 1-km MODIS LST pixels. (<b>a</b>), (<b>b</b>) and (<b>c</b>) respectively show the monthly RB of the stations in the upstream, midstream and downstream of the Heihe River Basin (HRB).</p>
Full article ">Figure 6
<p>Monthly normalized difference vegetation index (NDVI) at the eighteen sites from 2013 to 2014.</p>
Full article ">Figure 7
<p>Monthly average structure scale (ASS) of the stations. (<b>a</b>), (<b>b</b>) and (<b>c</b>) respectively show the monthly ASS of the stations in the upstream, midstream and downstream of the HRB.</p>
Full article ">Figure 8
<p>Grading results of the stations. (<b>a</b>,<b>b</b>) show the monthly representativeness level of the upstream stations; (<b>c</b>,<b>d</b>) show the monthly representativeness level of midstream and downstream stations</p>
Full article ">Figure 9
<p>Scatterplots of the daytime comparison results of the LSTs from MODIS and the LST measurements at the eighteen HiWATER sites.</p>
Full article ">Figure 10
<p>Scatterplots of the nighttime comparison results of the LSTs from MODIS and the LST measurements at the eighteen HiWATER sites.</p>
Full article ">Figure 11
<p>Relationships between the bias and zenith angle of MODIS when viewing the pixels: (<b>a</b>) relationship during the daytime; and (<b>b</b>) relationship at nighttime.</p>
Full article ">
4914 KiB  
Article
Estimating Leaf Area Density of Individual Trees Using the Point Cloud Segmentation of Terrestrial LiDAR Data and a Voxel-Based Model
by Shihua Li, Leiyu Dai, Hongshu Wang, Yong Wang, Ze He and Sen Lin
Remote Sens. 2017, 9(11), 1202; https://doi.org/10.3390/rs9111202 - 22 Nov 2017
Cited by 91 | Viewed by 11535
Abstract
The leaf area density (LAD) within a tree canopy is very important for the understanding and modeling of photosynthetic studies of the tree. Terrestrial light detection and ranging (LiDAR) has been applied to obtain the three-dimensional structural properties of vegetation and estimate the [...] Read more.
The leaf area density (LAD) within a tree canopy is very important for the understanding and modeling of photosynthetic studies of the tree. Terrestrial light detection and ranging (LiDAR) has been applied to obtain the three-dimensional structural properties of vegetation and estimate the LAD. However, there is concern about the efficiency of available approaches. Thus, the objective of this study was to develop an effective means for the LAD estimation of the canopy of individual magnolia trees using high-resolution terrestrial LiDAR data. The normal difference method based on the differences in the structures of the leaf and non-leaf components of trees was proposed and used to segment leaf point clouds. The vertical LAD profiles were estimated using the voxel-based canopy profiling (VCP) model. The influence of voxel size on the LAD estimation was analyzed. The leaf point cloud’s extraction accuracy for two magnolia trees was 86.53% and 84.63%, respectively. Compared with the ground measured leaf area index (LAI), the retrieved accuracy was 99.9% and 90.7%, respectively. The LAD (as well as LAI) was highly sensitive to the voxel size. The spatial resolution of point clouds should be the appropriate estimator for the voxel size in the VCP model. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Field measurements. (<b>a</b>) Magnolia A; (<b>b</b>) Magnolia B; (<b>c</b>) The location of scanning stations and reference targets, with the dots representing reference targets that are used to establish the correspondences between different scanning stations (the squares); light detection and ranging LiDAR point clouds of Magnolia A (<b>d</b>); and Magnolia B (<b>e</b>); (<b>f</b>) LiDAR point clouds of leaves.</p>
Full article ">Figure 2
<p>Flowchart of leaf area density (LAD) estimation.</p>
Full article ">Figure 3
<p>Normals of a point cloud of the leaf and trunk or branch. The red dot is point p, the gray dots are the other points in the neighborhood (yellow sphere). The arrows are the normals of these points.</p>
Full article ">Figure 4
<p>The schematic diagram of a Voxel-based model. (<b>a</b>) Illustration of the canopy voxelization; (<b>b</b>) the interception (1) and non-interception (0) of the laser beam within a horizontal layer; and (<b>c</b>) the vertical distribution of intercepted voxels.</p>
Full article ">Figure 5
<p>Leaf area index (LAI) measurement of a tree using LAI 2200™ sensor [<a href="#B36-remotesensing-09-01202" class="html-bibr">36</a>]. The sensor is placed near the base of a trunk with a 90° view cap.</p>
Full article ">Figure 6
<p>The leaf point clouds extraction of Magnolia A. (<b>a</b>) The point clouds of the whole canopy. The yellow rectangle is the tested cube; (<b>b</b>) the point clouds of test cube 1; (<b>c</b>) the point clouds of leaves; (<b>d</b>) the segmented leaves of Magnolia A; (<b>e</b>) the segmented leaves of test cube 1; (<b>f</b>) the segmented leaves.</p>
Full article ">Figure 7
<p>The leaf point clouds extraction of Magnolia B. (<b>a</b>) The point clouds of the whole canopy. The yellow rectangle is the tested cube; (<b>b</b>) the point clouds of test cube 4; (<b>c</b>) the point clouds of leaves; (<b>d</b>) the segmented leaves of the tree; (<b>e</b>) the segmented leaves of test cube 4; (<b>f</b>) the segmented leaves.</p>
Full article ">Figure 8
<p>Outline of the horizontal layer. Each small black point represents the interception of the laser beam at a voxel location. A large red dot denotes the convex hull polygon vertex of the point sets, that is, the boundary layer of the tree canopy vertices.</p>
Full article ">Figure 9
<p>Probability distribution of the leaf inclination angle of Magnolia A.</p>
Full article ">Figure 10
<p>Probability distribution of the leaf inclination angle of Magnolia B.</p>
Full article ">Figure 11
<p>Leaf area density (LAD) profile of the individual broadleaf trees.</p>
Full article ">Figure 12
<p>Variation for the contact frequency with the voxel size in different horizontal thickness layers of Magnolia A. (<b>a</b>–<b>i</b>) indicate layers 1 (the lowest) to 9 (the highest), respectively.</p>
Full article ">Figure 13
<p>LAI of magnolia trees for different voxel sizes. (<b>a</b>) Magnolia A; (<b>b</b>) Magnolia B.</p>
Full article ">
8274 KiB  
Article
Advancing the PROSPECT-5 Model to Simulate the Spectral Reflectance of Copper-Stressed Leaves
by Chengye Zhang, Huazhong Ren, Yanzhen Liang, Suhong Liu, Qiming Qin and Okan K. Ersoy
Remote Sens. 2017, 9(11), 1191; https://doi.org/10.3390/rs9111191 - 20 Nov 2017
Cited by 14 | Viewed by 5443
Abstract
This paper proposes a modified model based on the PROSPECT-5 model to simulate the spectral reflectance of copper-stressed leaves. Compared with PROSPECT-5, the modified model adds the copper content of leaves as one of input variables, and the specific absorption coefficient related to [...] Read more.
This paper proposes a modified model based on the PROSPECT-5 model to simulate the spectral reflectance of copper-stressed leaves. Compared with PROSPECT-5, the modified model adds the copper content of leaves as one of input variables, and the specific absorption coefficient related to copper (Kcu) was estimated and fixed in the modified model. The specific absorption coefficients of other biochemical components (chlorophyll, carotenoid, water, dry matter) were the same as those in PROSPECT-5. Firstly, based on PROSPECT-5, we estimated the leaf structure parameters (N), using biochemical contents (chlorophyll, carotenoid, water, and dry matter) and the spectra of all the copper-stressed leaves (samples). Secondly, the specific absorption coefficient related to copper (Kcu) was estimated by fitting the simulated spectra to the measured spectra using 22 samples. Thirdly, other samples were used to verify the effectiveness of the modified model. The spectra with the new model are closer to the measured spectra when compared to that with PROSPECT-5. Moreover, for all the datasets used for validation and calibration, the root mean square errors (RMSEs) from the new model are less than that from PROSPECT-5. The differences between simulated reflectance and measured reflectance at key wavelengths with the new model are nearer to zero than those with the PROSPECT-5 model. This study demonstrated that the modified model could get more accurate spectral reflectance from copper-stressed leaves when compared with PROSPECT-5, and would provide theoretical support for monitoring the vegetation stressed by copper using remote sensing. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The experimental scene of some samples: (<b>a</b>) The distribution of plastic pots of wheat. (The pots distribution of pak choi is same with that of wheat. Circle: a pot); (<b>b</b>) Some samples of wheat (stress level: 400 mg/kg; 800 mg/kg); (<b>c</b>) Some samples of pak choi (stress level: 0 mg/kg; 25 mg/kg).</p>
Full article ">Figure 2
<p>The steps of the method used in this study.</p>
Full article ">Figure 3
<p>Scanning electron microscopy (SEM). images of leaves with different copper contents in soil: (<b>a</b>) Normal wheat; (<b>b</b>) Wheat with 200 mg/kg copper content in soil; (<b>c</b>) Wheat with 400 mg/kg copper content in soil; (<b>d</b>) Wheat with 1600 mg/kg copper content in soil; (<b>e</b>) Normal pak choi; (<b>f</b>) Pak choi with 200 mg/kg copper content in soil; (<b>g</b>) Pak choi with 400 mg/kg copper content in soil; (<b>h</b>) Pak choi with 800 mg/kg copper content in soil.</p>
Full article ">Figure 4
<p>The modeled reflectance generated by PROSPECT-5 with different <span class="html-italic">N</span>. (<span class="html-italic">C<sub>ab</sub></span>, <span class="html-italic">C<sub>car</sub></span>, <span class="html-italic">C<sub>w</sub></span>, and <span class="html-italic">C<sub>m</sub></span> are fixed as 33 μg/cm<sup>2</sup>, 8.6 μg/cm<sup>2</sup>, 0.012 cm, 0.005 g/cm<sup>2</sup>, respectively).</p>
Full article ">Figure 5
<p>Leaf structure parameters <span class="html-italic">N</span> of copper-stressed leaves and normal leaves: (<b>a</b>) The value of leaf structure parameters; (<b>b</b>) The box-and-whisker plots of leaf structure parameters; (<b>c</b>) Leaf structure parameters vs. leaf copper content with different symbols for each species; (<b>d</b>) Bar graphs for the average of different stress levels (low, medium, high) and species.</p>
Full article ">Figure 6
<p>The specific absorption coefficient related to copper (<span class="html-italic">K<sub>cu</sub></span>).</p>
Full article ">Figure 7
<p>The simulated spectra (PROSPECT-5 and the new model) and measured spectra: (<b>a</b>) The mean (±standard deviation, SD) of simulated spectra and measured spectra of the 11 samples; (<b>b</b>) The selected representative sample: No. 5 sample (wheat: <span class="html-italic">C<sub>ab</sub></span> = 49.63 μg/cm<sup>2</sup>; <span class="html-italic">C<sub>car</sub></span> = 8.94 μg/cm<sup>2</sup>; <span class="html-italic">C<sub>w</sub></span> = 0.0096 cm; <span class="html-italic">C<sub>m</sub></span> = 0.0061 g/cm<sup>2</sup>; <span class="html-italic">C<sub>cu</sub></span> = 0.0557 μg/cm<sup>2</sup>); (<b>c</b>) The selected representative sample: No. 7 sample (pak choi: <span class="html-italic">C<sub>ab</sub></span> = 27.61 μg/cm<sup>2</sup>; <span class="html-italic">C<sub>car</sub></span> = 19.32 μg/cm<sup>2</sup>; <span class="html-italic">C<sub>w</sub></span> = 0.0320 cm; <span class="html-italic">C<sub>m</sub></span> = 0.0031 g/cm<sup>2</sup>; <span class="html-italic">C<sub>cu</sub></span> = 0.1520 μg/cm<sup>2</sup>).</p>
Full article ">Figure 8
<p>Differences between simulated reflectance and measured reflectance at key wavelengths.</p>
Full article ">
11340 KiB  
Article
Satellite-Derived Spatiotemporal Variations in Evapotranspiration over Northeast China during 1982–2010
by Lilin Zhang, Yunjun Yao, Zhiqiang Wang, Kun Jia, Xiaotong Zhang, Yuhu Zhang, Xuanyu Wang, Jia Xu and Xiaowei Chen
Remote Sens. 2017, 9(11), 1140; https://doi.org/10.3390/rs9111140 - 7 Nov 2017
Cited by 15 | Viewed by 4286
Abstract
Evapotranspiration (ET) is a critical process for the climate system and water cycles. However, the spatiotemporal variations in terrestrial ET over Northeast China over the past three decades calculated from sparse meteorological point-based data remain large uncertain. In this paper, a recently proposed [...] Read more.
Evapotranspiration (ET) is a critical process for the climate system and water cycles. However, the spatiotemporal variations in terrestrial ET over Northeast China over the past three decades calculated from sparse meteorological point-based data remain large uncertain. In this paper, a recently proposed modified satellite-based Priestley–Taylor (MS–PT) algorithm was applied to estimate ET of Northeast China during 1982–2010. Validation results show that the square of the correlation coefficients (R2) for the six flux tower sites varies from 0.55 to 0.88 (p < 0.01), and the mean root mean square error (RMSE) is 0.92 mm/d. The ET estimated by MS–PT has an annual mean of 441.14 ± 18 mm/year in Northeast China, with a decreasing trend from southeast coast to northwest inland. The ET also shows in both annual and seasonal linear trends over Northeast China during 1982–2010, although this trend seems to have ceased after 1998, which increased on average by 12.3 mm per decade pre-1998 (p < 0.1) and decreased with large interannual fluctuations post-1998. Importantly, our analysis on ET trends highlights a large difference from previous studies that the change of potential evapotranspiration (PET) plays a key role for the change of ET over Northeast China. Only in the western part of Northeast China does precipitation appear to be a major controlling influence on ET. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Map of the land cover type and the location of flux towers (Jinzhou, Duolun, Changbaishan, Tongyu, Laoshan) in Northeast China. The two main basins are also shown in inset panel: Songjiang (SJ) basin and Liaohe (LH) basin.</p>
Full article ">Figure 2
<p>The scatterplots between flux tower observations and estimated ET at daily scale: (<b>a</b>) Changbaishan; (<b>b</b>) Laoshan; (<b>c</b>) Duolun1; (<b>d</b>) Duolun2; (<b>e</b>) Jinzhou; (<b>f</b>) Tongyu. The bias and RMSE are in units of mm/d. The solid line is the 1:1 line.</p>
Full article ">Figure 3
<p>Comparisons of the predicted and measured site-averaged ET at six sites. The bias and RMSE are in units of mm/d. The solid line is the 1:1 line.</p>
Full article ">Figure 4
<p>The monthly time series comparison of MS–PT-based ET and MOD16 ET in the Changbaishan station. RMSE is in units of mm/month.</p>
Full article ">Figure 5
<p>Spatial patterns of (<b>a</b>) multiyear average ET, (<b>b</b>) multiyear average precipitation, and (<b>c</b>) multiyear average soil moisture in Northeast China. All are in units of mm/year.</p>
Full article ">Figure 6
<p>Multiyear seasonal patterns of ET in Northeast China: (<b>a</b>) MAM (March, April, and May); (<b>b</b>) JJA (June, July, and August); (<b>c</b>) SON (September, October, and November); (<b>d</b>) DJF (December, January, and February). The ET are in units of mm/month.</p>
Full article ">Figure 7
<p>(<b>a</b>) Interannual variability of ET from 1982 to 2010 in Northeast China; partly interannual variability of ET: (<b>b</b>) the LH basin; and (<b>c</b>) the SJ basin.</p>
Full article ">Figure 8
<p>First column: interannual variability of ET from 1982–2010 in Northeast China: (<b>a</b>) MAM; (<b>b</b>) JJA; (<b>c</b>) SON; (<b>d</b>) DJF. Second column: interannual variability of ET in the SJ basin: (<b>a1</b>) MAM; (<b>b1</b>) JJA; (<b>c1</b>) SON; (<b>d1</b>) DJF. Third column: interannual variability of ET in the LH basin: (<b>a2</b>) MAM; (<b>b2</b>) JJA; (<b>c2</b>) SON; (<b>d2</b>) DJF.</p>
Full article ">Figure 9
<p>Spatial distributions of ET trends in Northeast China: (<b>a</b>) during 1982–2010; (<b>b</b>) during 1982–1998; (<b>c</b>) during 1998–2010. The ET trend is in units of mm/year. The inset panels show the area where the ET trend is statistically significant (<span class="html-italic">p</span> &lt; 0.05). Red represents a significant increase and blue represents a significant decrease.</p>
Full article ">Figure 10
<p>(First column) Spatial distributions of ET trends in Northeast China during 1982–2010: (<b>a</b>) MAM; (<b>b</b>) JJA; (<b>c</b>) SON; (<b>d</b>) DJF; (second column) during 1982–1998: (<b>a1</b>) MAM; (<b>b1</b>) JJA; (<b>c1</b>) SON; (<b>d1</b>) DJF; (third column) during 1998–2010: (<b>a2</b>) MAM; (<b>b2</b>) JJA; (<b>c2</b>) SON; (<b>d2</b>) DJF. The inset panels show the area where the ET trend is statistically significant (<span class="html-italic">p</span> &lt; 0.1).</p>
Full article ">Figure 10 Cont.
<p>(First column) Spatial distributions of ET trends in Northeast China during 1982–2010: (<b>a</b>) MAM; (<b>b</b>) JJA; (<b>c</b>) SON; (<b>d</b>) DJF; (second column) during 1982–1998: (<b>a1</b>) MAM; (<b>b1</b>) JJA; (<b>c1</b>) SON; (<b>d1</b>) DJF; (third column) during 1998–2010: (<b>a2</b>) MAM; (<b>b2</b>) JJA; (<b>c2</b>) SON; (<b>d2</b>) DJF. The inset panels show the area where the ET trend is statistically significant (<span class="html-italic">p</span> &lt; 0.1).</p>
Full article ">Figure 11
<p>Map of the correlation coefficient between the annual ET trend (1982–2010) and the trend of (<b>a</b>) precipitation, (<b>b</b>) soil moisture, (<b>c</b>) Tair, and (<b>d</b>) PET. The inset panels show the frequency of various change trends. The degree of correlation is classified into three ranks according to the correlation coefficient: strong (R<sup>2</sup> ≥ 0.7); moderate (0.3 &lt; R<sup>2</sup> &lt; 0.7); weak (R<sup>2</sup> ≤ 0.3).</p>
Full article ">

Review

Jump to: Editorial, Research

30 pages, 5292 KiB  
Review
Characterizing Land Surface Anisotropic Reflectance over Rugged Terrain: A Review of Concepts and Recent Developments
by Jianguang Wen, Qiang Liu, Qing Xiao, Qinhuo Liu, Dongqin You, Dalei Hao, Shengbiao Wu and Xingwen Lin
Remote Sens. 2018, 10(3), 370; https://doi.org/10.3390/rs10030370 - 27 Feb 2018
Cited by 120 | Viewed by 10747
Abstract
Rugged terrain, including mountains, hills, and some high lands are typical land surfaces around the world. As a physical parameter for characterizing the anisotropic reflectance of the land surface, the importance of the bidirectional reflectance distribution function (BRDF) has been gradually recognized in [...] Read more.
Rugged terrain, including mountains, hills, and some high lands are typical land surfaces around the world. As a physical parameter for characterizing the anisotropic reflectance of the land surface, the importance of the bidirectional reflectance distribution function (BRDF) has been gradually recognized in the remote sensing community, and great efforts have been dedicated to build BRDF models over various terrain types. However, on rugged terrain, the topography intensely affects the shape and magnitude of the BRDF and creates challenges in modeling the BRDF. In this paper, after a brief introduction of the theoretical background of the BRDF over rugged terrain, the status of estimating land surface BRDF properties over rugged terrain is comprehensively reviewed from a historical perspective and summarized in two categories: BRDFs describing solo slopes and those describing composite slopes. The discussion focuses on land surface reflectance retrieval over mountainous areas, the difference in solo slope and composite slope BRDF models, and suggested future research to improve the accuracy of BRDFs derived with remote sensing satellites. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Literature statistics for bidirectional reflectance distribution function (BRDF) modeling over rugged terrain contributed by different research field in recent decades.</p>
Full article ">Figure 2
<p>Literature statistics for BRDF modeling over rugged terrain from 1983 to 2017. (<b>a</b>) The numbers of published articles, and (<b>b</b>) total citations.</p>
Full article ">Figure 3
<p>Configuration of solar illumination and sensor over a slope surface.</p>
Full article ">Figure 4
<p>Graphics of topography relief: (<b>a</b>) nature surface of solo slope, (<b>b</b>) the topographic model of solo slope, (<b>c</b>) nature of composite slope, and (<b>d</b>) topographic model of composite slope.</p>
Full article ">Figure 5
<p>Reference plane configuration over solo slope ((<b>a</b>) slope-parallel white plane and (<b>b</b>) horizontal reference plane) and composite slope ((<b>c</b>) horizontal reference plane at the highest point).</p>
Full article ">Figure 6
<p>Key procedures of BRDF modeling over rugged terrain.</p>
Full article ">Figure 7
<p>Irradiance at the land surface.</p>
Full article ">Figure 8
<p>Canopy shadow cast on flat and sloped forest. (<b>a</b>) Flat forest. (<b>b</b>,<b>c</b>) Sloped forest. The dotted lines represent the incident solar beam.</p>
Full article ">Figure 9
<p>Topographic effects on crown sun-canopy-sensor geometry. (<b>a</b>) Forest stand on solo slope surface, (<b>b</b>) geometry correction without negative geotropism consideration, and (<b>c</b>) geometry correction with negative geotropism consideration.</p>
Full article ">Figure 10
<p>Solo slope reflectance simulated by the GOMST extended by the SAIL model and coupled topography (GOSAILT) model, where (<b>a</b>–<b>c</b>) are the red reflectance and the (<b>d</b>–<b>f</b>) are the NIR reflectance. The solar zenith is 30° and azimuth is 0°. The slopes aspect are also 0°; (<b>a</b>,<b>d</b>) are the flat terrain; (<b>b</b>, <b>e</b>) are the 30° slope; and (<b>c</b>,<b>f</b>) are the 60° slope; Red lines indicate the BRFs along the PP. The radial distance and polar angle of polar coordinate system are view zenith angle and the view azimuth angle, respectively.</p>
Full article ">Figure 11
<p>Radiative transfer process over the composite slope terrain.</p>
Full article ">Figure 12
<p>Modeled surfaces with different spherical shape hypotheses.</p>
Full article ">Figure 13
<p>Random surface with normal distribution.</p>
Full article ">Figure 14
<p>Equivalent slope: a virtual smooth surface.</p>
Full article ">Figure 15
<p>Global topographic shadow mask (TSM).</p>
Full article ">
Back to TopTop